Sie sind auf Seite 1von 22

Appendix: Evolution of Hydraulic

Fracturing Design and Evaluation


K. G. Nolte, Schlumberger Dowell

Overview
This Appendix to Chapter 5 reviews the evolution
of hydraulic fracturing design and evaluation
methods. Complementary reviews are the application of fracturing by Smith and Hannah (1996) and
fracturing fluids by Jennings (1996). This review of
design and evaluation considers three generations of
fracturing: damage bypass, massive treatments and
tip-screenout (TSO) treatments.
The first two generations of fracturing and their
links to practices are emphasized because these contributions are not likely well known by the current
generation of engineers. The review focuses on
propped fracturing and does not explicitly consider
acid fracturing. Although the principles governing
the mechanics of both are essentially the same, the
fluid chemistry for obtaining fracture conductivity
is quite different (see Chapter 7). These principles
have their roots in civil and mechanical engineering,
more specifically in the general area of applied
mechanics: solid mechanics for the rock deformation and fluid mechanics for the flow within the
fracture and porous media. For the porous media
aspects, fracturing evaluation has benefited greatly
from the reservoir engineering practices discussed
in Chapters 2 and 12.
This review reflects the authors perspective and
bias in interpreting the impact of past contributions,
and therefore parts of this review should be anticipated to raise objections from others with an extensive knowledge of fracturing. In addition to this
volume, the Society of Petroleum Engineers (SPE)
Monograph Recent Advances in Hydraulic Fracturing (Gidley et al., 1989) provides balanced,
detailed coverage of the diverse areas of fracturing
from the perspectives of more than 30 fracturing
specialists.
This review concludes with speculation concerning a future generation, in which fracture design and
reservoir engineering merge into fracturing for

Reservoir Stimulation

reservoir management (i.e., control of both the vertical and horizontal flow profiles within the reservoir). Similar speculation in a 1985 lecture suggested that development of the technical foundation
for the TSO generation would quickly bring higher
permeability formations into consideration as typical
fracturing candidates (i.e., moderate k (2) on
Appendix Fig. 1a, with 2 indicating a target for
folds of increase [FOI] in the production rate, in
contrast to 10 for tight gas and massive treatments). However, the advent of this generation was
considerably delayed because of two factors that
have generally dominated technical considerations
during the history of fracturing. These dominating
factors are hydrocarbon prices and resistance to trying something new until established practices fail to
allow the economic development of a prospect.
The cycles of fracturing activity in Appendix Fig.
1a clearly reflect the timing of the first two fracturing generations. Appendix Fig. 1b identifies economic drivers for corresponding cycles in the U.S.
rig count. The first surge of activity resulted when
rotary drilling was introduced, which enabled the
development of deeper reserves. Fracturing activity
followed this trend soon after its commercialization
in 1949 because it was found to be an effective,
low-cost means of mitigating the resulting drilling
mud damage to reservoir sections (i.e., the damage
bypass generation). Both drilling and fracturing
activities began a long-term decline after 1955
because of degrading prices caused by imported oil
and regulated gas prices. Similarly, both activities
began a rapid increase at about 1979 as prices
increased because the Organization of Petroleum
Exporting Countries (OPEC) reduced its oil supplies
and a natural gas shortage developed in the United
States. The gas shortage, and its 10-fold-plus
increase in price, encouraged the development of
tight gas reserves and an associated demand for
massive fracturing treatments to develop the tight
reserves. The failure of past fracturing practices for

A5-1

(a)

Mode
rate k
(2)
High c
onduc
tivity

Un
T
de ight
rst
ga
an
s(
eq ding 10)
uip , m
me ate
nt
ria
ls,

Rem
ove
dam
age

Treatments

4000

0
1950 1955

1971
Year

1981

15.1

Up

Up

9.
9%

/y
ea
r

%/y

ear

4500
OPEC overextends
4000
Prices fall
U.S. gas prices regulated
Middle
East
discoveries
3500
3000
U.S. production peaks
OPEC develops
2500
price authority
Do
wn
2000
6.
1%
/ye
1500
ar
1000
Rotary displaces
500
cable tool drilling
0
1940 1950 1960 1970 1980

1985 forcast flat

.5%/yea

Down 25

The beginning

Annual average rotary rig count

(b)

1990

2000

Year

Appendix Figure 1. (a) Trends in fracturing activity treatments per month (courtesy of K. G. Nolte and M. B. Smith,
19851986 SPE Distinguished Lecture). (b) U.S. drilling rig
activity shows five major trends (updated from Oil & Gas
Journal, 1985).

large treatments spurred a significant research and


development effort that beneficially impacted every
aspect of fracturing and essentially developed the
fracture design and evaluation framework presented
in this volume. The industrys rapid contraction during the early 1980s resulted again from OPEC, but
this time because of OPECs failure to maintain artificially high prices. The TSO treatment for creating
the very wide propped fractures required for high
permeability evolved during this time. This technique allowed the development of a troublesome
soft-chalk reservoir in the North Sea by fracturing.
However, the significant potential of the TSO generation did not materialize until about 10 years later,
when its application was required on a relatively
large scale to achieve viable economics for two highpermeability applications: bypassing deep damage in
the Prudhoe Bay field and its coupling with gravel

A5-2

packing to achieve low-skin completions for a significant venture in the Gulf of Mexico.
The potential for a future reservoir management
generation was demonstrated in 1994 for the Norwegian Gullfaks field. The potential is to use TSO
treatments and indirect vertical fracturing for
increased reserves recovery, formation solids control
and water management. However, the unique benefits and favorable economics for this different
approach to reservoir plumbing were slow to
materialize because of the industrys comfort with
deviated drilling and more traditional completions.
Another observation from this historical perspective is the 1985 forecast of a flat drilling level
(Appendix Fig. 1b). However, activity continued to
decrease rapidly, to less than one-half of the forecast,
and subsequently declined by another one-half. Stable
activity levels within the petroleum industry are not
seen in the historical cycles and remain the product
of wishful thinking.

The concept of hydraulic fracturing within the petroleum industry was developed during the last half of
the 1940s within Stanolind (now BP Amoco; e.g.,
Clark, 1949; Farris, 1953; Howard and Fasts
Hydraulic Fracturing Monograph, 1970) by building
on the industrys experience with injection techniques that had experienced increased injectivity
by fracturing: acidizing (Grebe and Stoesser, 1935),
squeeze cementing and brine injection wells. A reissued patent was granted (Farris, 1953, resulting
from an initial filing in 1948) that was comprehensive in scope and covered many recognized practices
and products: proppant, gelled oil, breakers, fluidloss additives, continuous mixing, pad-acid fracturing, emulsified acids and the use of packers for fracturing multiple zones. Several aspects of the patent
that later became important included the implication
that fractures were horizontal and the use of a lowpenetrating fluid or with viscosity > 30 cp.
The first experimental treatments were performed
in 1947 on four carbonate zones in the Houghton
field in Kansas (Howard and Fast, 1970). The zones
had been previously acidized and were isolated by a
cup-type straddle packer as each was treated with
1000 gal of napalm-thickened gasoline followed by
2000 gal of gasoline as a breaker. These unpropped

Chapter 5 Appendix: Evolution of Hydraulic Fracturing Design and Evaluation

treatments did not increase production and led to the


incorrect belief for some time that fracturing had no
benefit over acidizing for carbonate formations.
A subsequent treatment of the Woodbine sand in
the East Texas field was highly successful. It consisted of 23 bbl of gelled lease crude, 160 lbm of
16-mesh sand at 0.15 ppa and 24 bbl of breaker
(Farris, 1953). Halliburton originally obtained an
exclusive license from Stanolind and commercialized fracturing in 1949. Activity rapidly expanded to
about 3000 treatments per month by 1955 (Appendix
Fig. 1a). Before a universal license was granted to
other service companies, water or river fracturing
became popular in lower permeability areas such as
the San Juan basin (C. R. Fast, pers. comm., 1997).
As implied by the name, the treatments used river
water and sand. The water was outside the definition
of a nonpenetrating fluid within the patents specified
filtrate rate through filter paper or viscosity greater
than 30 cp.

The first generation: damage bypass


Applications of first-generation fracturing were primarily small treatments to bypass near-wellbore
drilling fluid damage to formations with permeability
in the millidarcy range. An inherent advantage of
propped fracturing, relative to matrix treatment for
damage removal, is that a fracture opens the complete section and retains a conductive path into the
zone. The complete opening overcomes the diversion
consideration for matrix treatments (see Chapter 19),
but adds the consideration of producing from bottomwater or an upper gas cap. For lower permeability
formations, large amounts of produced water are
generally not a problem. For higher permeability formations, water production can be significant, which
provided the historical preference for matrix treatment in higher permeability applications. However,
the precision of fracturing improved significantly,
and TSO treatments have been routinely performed
in Prudhoe Bay oil columns only 50 ft thick and
above very mobile water (Martins et al., 1992b).
The technology for this fracturing generation is
summarized in the Howard and Fast (1970) Monograph. The breadth of this volume is shown by its
comprehensive consideration of candidate selection
(see Chapter 1) and optimal design based on economic return (see Chapters 5 and 10). Other note-

Reservoir Stimulation

worthy design and evaluation methods from this generation are fracture orientation (horizontal or vertical), in-situ stress and fracture width models, FOI
prediction and fracture conductivity in production
enhancement.

Fracture orientation and in-situ stress


The application of mechanics to fracturing was catalyzed by the horizontal orientation of fractures
implied in the Stanolind patent and the desire of several operators to avoid paying the nominal patent
royalty of $25$125, based on volume (C. R. Fast,
pers. comm., 1997). Significant research activity was
conducted to show that fractures can be vertical, as is
now known to be the general case for typical fracturing conditions. The fracture orientation debate eventually led to a lawsuit that was settled before the trial
ended. The settlement accepted the patent and nominal royalty payments and stipulated that other service
companies receive a license to practice fracturing.
However, the royalty benefits were more than nominal to Stanolind because about 500,000 treatments
were performed during the 17-year period of the
patent (C. R. Fast, pers. comm., 1997). Key to
the favorable settlement for Stanolind was its welldocumented demonstration of a horizontal fracture
in the Pine Island field (see fig. 7-1 in Howard and
Fast, 1970).
The central issue in the orientation debate was the
direction of the minimum stress. The pressure
required to extend a fracture must exceed the stress
acting to close the fracture. Therefore, the fracture
preferentially aligns itself perpendicular to the direction of minimum stress because this orientation provides the lowest level of power to propagate the fracture. The minimum stress direction is generally horizontal; hence, the fracture plane orientation is generally vertical (i.e., a vertical fracture). The preference
for a horizontal fracture requires a vertical minimum
stress direction.
In the following review, the orientation consideration is expanded to also cover the state of stress in
more general terms. The stress at any point in the various rock layers intersected by the fracture is defined
by its magnitude in three principal and perpendicular
directions. The stress state defines not only the fracture orientation, but also the fluid pressure required to
propagate a fracture that has operational importance,
vertical fracture growth into surrounding formation

A5-3

h = Ko v ,

(1)

where Ko = /(1 ) = 13 for = 14 (see Eq. 3-51).


Using Poissons ratio of 14, Harrison et al. concluded that the horizontal stress is about one-third of
the vertical stress and therefore fractures are vertical.
Appendix Eq. 1 provides the current basis for using
mechanical properties logs to infer horizontal stress,
with Poissons ratio obtained from a relation based
on the shear and compressional sonic wave speeds
(see Chapter 4). Another assumption for Appendix
Eq. 1 is uniaxial compaction, based on the premise
that the circumference of the earth does not change
as sediments are buried to the depths of petroleum
reservoirs and hence the horizontal components of
strain are zero during this process. Therefore, Appendix Eq. 1 provides the horizontal stress response
to maintain the horizontal dimensions of a unit cube
constant under the application of vertical stress.
However, there is one problem with this 1954
conclusion concerning horizontal stress. Appendix
Eq. 1 is correct for the effective stress but not for
the total stress that governs fracture propagation:
= p, where p is the pore pressure, which also
has a role in transferring the vertical stress into horizontal stress as explicitly shown by Appendix Eq. 2.
Harrison et al. (1954) correctly postulated that shales
have higher horizontal stresses and limit the vertical
fracture height. The general case of higher stress in
shales than in reservoir rocks was a necessary condition for the successful application of fracturing
because fractures follow the path of least stress. If
the converse were the general case, fractures would
prefer to propagate in shales and not in reservoir
zones.

A5-4

Harrison et al. also reported the Sneddon and


Elliott (1946) width relation for an infinitely extending pressurized slit contained in an infinitely extending elastic material. This framework has become the
basis for predicting fracture width and fracturing
pressure response (see Chapters 5, 6 and 9). They
used the fracture length for the characteristic, or
smaller and finite, dimension in this relation. Selecting the length for the characteristic dimension
resulted in what is now commonly termed the KGD
model. Selecting the height, as is the case for a very
long fracture, is termed the PKN model. These models are discussed in the next section and Chapter 6.
Harrison et al. considered a width relation because
of its role in fracture design to determine the fluid
volume required for a desired fracture extent.
The role of volume balance (or equivalently, the
material balance in reservoir terminology) is an
essential part of fracture design and fracture simulation code. As shown schematically on the left side of
Appendix Fig. 2, each unit of fluid injected Vi is
either stored in the fracture to create fracture volume
or lost to the formation as fluid loss. (However,
Harrison et al.s 1954 paper does not discuss fluid
loss.) The stored volume is the product of twice the
fracture half-length L, height hf and width w. If the
latter two dimensions are not constant along the fracture length, they can be appropriately averaged over
the length. The half-length is then obtained by simply dividing the remaining volume, after removing
the fluid-loss volume, by twice the product of the

L=

Vi
2hf (w + CL 8t )
=

2hf wL
Vi

Fluid loss
CL t
Proppant

layers and stress acting to crush proppant or to close


etched channels from acid fracturing. The crushing
stress is the minimum stress minus the bottomhole
flowing pressure in the fracture. The orientation
debate resulted in three papers that will remain significant well into the future.
The first paper to be considered is by Harrison et
al. (1954). Some of the important points in the paper
are that the overburden stress (vertical stress v) is
about 1 psi per foot of depth, fracturing pressures are
generally lower than this value and therefore fractures are not horizontal, and an inference from elasticity that the minimum horizontal stress is

Geometry
hf

2L

Proppant
(% area = )

w
Pad

Volume

Appendix Figure 2. Volume balance for fracture placement (equation from Harrington et al., 1973) (adapted
courtesy of K. G. Nolte and M. B. Smith, 19841985 SPE
Distinguished Lecture).

Chapter 5 Appendix: Evolution of Hydraulic Fracturing Design and Evaluation

average height and the average width. The fluid-loss


volume depends on the fluid-loss surface area, or a
height-length product. Furthermore, as shown on the
right side of Appendix Fig. 2, the ratio of stored to
total volume is termed the fluid efficiency and directly affects the proppant additional schedule (Harrington et al., 1973; Nolte, 1986b) (see Sidebar 6L).
The second paper to be discussed from the orientation era is by Hubbert and Willis (1957). The lessons
from this paper extend beyond fracturing and into
the area of structural geology. This work provides
simple and insightful experiments to define the state
of in-situ stress and demonstrate a fractures preference to propagate in the plane with minimum stress
resistance. For the latter experiments, the formation was gelatin within a plastic bottle preferentially
stressed to create various planes of minimal stress.
They also used simple sandbox experiments to
demonstrate normal and thrust faulting and to define
the state of stress for these conditions (see Sidebar
3A). They showed that Ko, or equivalently the horizontal stress, within Appendix Eq. 1 is defined by
the internal friction angle ( = 30 for sand) and is
1
3 for the minimum stress during normal faulting and
3 for the maximum stress during thrust faulting. For
the normal faulting case and correctly including pore
pressure in Appendix Eq. 1, the total minimum horizontal stress becomes
h = ( v + 2 p) 3 ,

(2)

where Ko = 13 with = 30. For this case the horizontal stress is much less than the vertical stress except
in the extreme geopressure case of pore pressure
approaching overburden, which causes all stresses
and pore pressure to converge to the overburden
stress. For the thrust faulting case, the larger horizontal stress (i.e., for the two horizontal directions) is
greater than the overburden and the smaller horizontal stress is equal to or greater than the overburden.
Both the extreme geopressure case and an active
thrust faulting regime can lead to either vertical or
horizontal fractures. The author has found Appendix
Eq. 2 to accurately predict the horizontal stress in tectonically relaxed sandstone formations ranging from
microdarcy to darcy permeability. The accuracy at
the high range is not surprising, as the formations
approach the unconsolidated sand in the sandbox
experiments. The accuracy obtained for microdarcypermeability sands is subsequently explained.

Reservoir Stimulation

Hubbert and Willis also provided an important set


of postulates: the rock stresses within the earth are
defined by rock failure from tectonic action and the
earth is in a continuous state of incipient faulting.
From this perspective, the stress is not governed
by the behavior of the intact rock matrix, but by an
active state of failure along discrete boundaries (e.g.,
by sand grains within fault boundaries, which
explains the application of Appendix Eq. 2 to microdarcy-permeability sandstones). This insightful conclusion about the role of failure is at the other
extreme of the behavior spectrum from the elastic
assumptions that Poissons ratio (Appendix Eq. 1)
governs the horizontal stress and that failure has no
effect on the stress. This extreme difference in the
assumptions for Appendix Eqs. 1 and 2 is often
overlooked because of the similar value of Ko = ~13
obtained in the case of a tectonically relaxed region
and Poissons ratio near 14. However, the role of elasticity becomes important in thrusting areas (see
Section 3-5.2) because of the difference in horizontal
stress resulting for layers with different values of
Youngs modulus (stiffness). More of the tectonic
action and higher levels of stress are supported by the
stiffer layers.
Additional considerations for horizontal stress outlined by Prats (1981) include the role of long-term
creep. Creep deformation allows relaxation of the
stress difference between the overburden and horizontal stresses, thereby enabling the horizontal stress
to increase toward the larger vertical stress governed
by the weight of the overburden. This effect is well
known for salt layers that readily creep and can collapse casing by transferring most of the larger overburden stress into horizontal stress. The role of stress
relaxation is an important mechanism for providing
favorable stress differences between relatively clean
sands governed by friction (i.e., Appendix Eq. 2) with
minimal creep and sediments with higher clay content. In the latter case, the clay supports some of the
intergranular stresses. The clay structure is prone to
creep that relaxes the in-situ stress differences and
increases the horizontal stress for a clay-rich formation.
Hence, both clay content and Poissons ratio produce the same effect on horizontal stress. Because
clay content also increases Poissons ratio, there is
a positive correlation of clay content (creep-induced
stress) to larger Poissons ratios (and elastic stress,
from Appendix Eq. 1) inferred from sonic velocities.
The implication of the correlation is that clay-rich

A5-5

formations can also have horizontal stresses greater


than those predicted by either Appendix Eq. 1 or 2,
which is consistent with the general requirement to
calibrate elastic-based stress profiles to higher levels
of stress (e.g., Nolte and Smith, 1981). The correlation of clay and Poissons ratio links the conclusions
of Hubbert and Willis and Prats that horizontal stress
is governed primarily by nonelastic effects and the
general correlation between the actual stress and
elastic/sonic-based stress profiles.
The third significant paper from this period is by
Lubinski (1954). He was a Stanolind researcher who
introduced the role that poroelasticity can have in
generating larger stresses during fracturing. (Poroelasticity could increase horizontal stress and lead
to horizontal fractures, as in the Stanolind patent.)
Lubinski presented poroelasticity within the context
of its analogy to thermoelasticity. His use of the thermal stress analogy facilitates understanding the poroelastic concept because thermal stresses are generally
more readily understood than pore stresses by engineers. The analogy provides that when pore pressure
is increased in an unrestrained volume of rock, the
rock will expand in the same manner as if the temperature is increased. Conversely, when the pore
pressure is lowered, the rock will contract as if the
temperature is lowered. When the rock is constrained, as in a reservoir, a localized region of pore
pressure change will induce stress changes: increasing stress within the region of increasing pore pressure (e.g., from fracturing fluid filtrate or water

injection) and decreasing stress within the region


of decreasing pore pressure (e.g., production). The
long-term impact of Lubinskis paper is that the
importance of poroelasticity increases as routine
fracturing applications continue their evolution to
higher permeability formations. This is apparent
from the thermal analogyas the area of expansion
increases the induced stresses also increase. For
poroelasticity, the area of significant transient change
in pore pressure increases as the permeability
increases (see Section 3-5.8).
Appendix Fig. 3 shows an example of significant
poroelasticity for a frac and pack treatment in a
1.5-darcy oil formation. The time line for the figure
begins with two injection sequences for a linear-gel
fluid and shows the pressure increasing to about
7500 psi and reaching the pressure limit for the operation. During the early part of the third injection
period, crosslinked fluid reaches the formation and
the pressure drops quickly to about 5600 psi (the
native fracturing pressure) and remains essentially
constant during the remainder of the injection.
The first two injections, without a crosslinked-fluid
filtrate (or filter cake) to effectively insulate the formation (as in the thermal analogy) from the increasing injection pressure, resulted in pore pressure
increases of significant magnitude and extent within
the formation. The pore pressure increase provides up
to a 1900-psi horizontal and poroelasticity stress
increase that extends the fracturing pressure beyond
the operational limit, leading to the shut-in for the

10,000

50

8000

Crosslinked gel

40

Injection
Step rate

Linear gel

6000

30

4000

20
Injection
Step rate

Minifracture

Propped fracture

2000

Injection rate (bbl/min)

Bottomhole pressure, BHP (psi)

Linear gel

10
BHP
Injection rate

0
0

0.5

1.0

2.0

13.0

13.5

0
14.0

Time (hr)

Appendix Figure 3. High-permeability frac and pack treatment (Gulrajani et al., 1997b).

A5-6

Chapter 5 Appendix: Evolution of Hydraulic Fracturing Design and Evaluation

second injection. This increase is about one-third of


the native stress. However, during the two subsequent
injections the insulating effect of the crosslinked
fluids internal cake and filtrate allows fracture extension within essentially the native stress state. The
pressure drop supported by the cake and filtrate is
about 1300 psi, as reflected by the rapid pressure
decrease after the third injection. This decrease
occurs because of the rapid closure and cessation of
fluid loss (that activated the pressure drop), which is
the same reason that surface pressure decreases at the
cessation of injection and loss of pipe friction. The
last injection for the proppant treatment is also of
interest because of the absence of a poroelastic effect
during the initial linear-gel injection. This observation
indicates that the insulating effect remained effective
from the prior injection of crosslinked fluid.
For a normally pressured and tectonically relaxed
area, the maximum increase in horizontal stress
before substantial fracture extension is about onethird of the native horizontal stress (Nolte, 1997),
as was found for the case shown in Appendix Fig.
3. Also, for any pore pressure condition in a relaxed
area, the stress increase will not cause the horizontal
stress to exceed the overburden (i.e., cause horizontal
fracturing). However, as the example shows, without
fluid-loss control, poroelasticity can significantly
increase the fracturing pressure and extend it beyond
operational limits for high-permeability reservoirs.

Width models
The first rigorous coupling of fluid flow and the elastic response of the formation was reported by
Khristianovich and Zheltov (1955). They used a twodimensional (2D) formulation based on a complex
variable analysis. Their formulation was equivalent
to the length becoming the characteristic, or smaller,
dimension and provides the initial K for the KGD
width model discussed later and in Chapter 6. In
addition to being the first paper to provide the coupling of fluid flow and rock interaction that is the
embodiment of the hydraulic fracturing process, the
paper also identified the role for a fluid lag region at
the fracture tip. This low-pressure region, beyond the
reach of fracturing fluid and filling with pore fluid,
has a large, negative net pressure and acts as a clamp
at the fracture tip. The fluid lags clamping effect
provides the natural means to lower the potentially

Reservoir Stimulation

large tip-region stresses to a level that can be accommodated by the in-situ condition. The presence of
the lag region has been demonstrated by field experiments at a depth of 1400 ft at the U.S. Department of
Energy (DOE) Nevada Test Site (Warpinski, 1985).
Appendix Fig. 4 compares the Khristianovich and
Zheltov analytical results for width and pressure to
the corresponding parameters from the Warpinski
field results. For the analytical results, decreasing
values of the complex variable angle 0 toward the
right side of the figure correspond to relatively
smaller lag regions and larger differences between
the minimum stress and pressure in the lag region
(i.e., generally deeper formations). The width profiles clearly show the clamping action at the tip, and
the field data appear to be represented by a 0 valve
of about /8 for the analytical cases. Also noteworthy of the experimental results is that tests 4 through
7 with water and test 9 with gel show similar behavior when test 4, which had a relatively low injection
rate, is ignored. Tests 10 and 11 were with a gelled
fluid and clearly show progressively different behavior from the preceding tests because of the altered tip
behavior resulting from prior gel injections and the
residual gel filter cakes that fill the fracture aperture
after closure. The cakes have the consistency of silicon rubber and functionally provide an analogous
sealing affect for subsequent tests.
The practical importance of the lag region cannot
be overemphasized. The extent of the region, which
is extremely small in comparison with commercialscale fractures, adjusts to the degree required to
essentially eliminate the role of the rocks fracture
resistance or toughness (e.g., see SCR Geomechanics
Group, 1993) and to isolate the fluid path from all
but the primary opening within the multitude of
cracks (process zone) forming ahead of the fracture
(see Chapters 3 and 6). The field data show the
width at the fluid front is well established (i.e., generally greater than 5% of the maximum width at the
wellbore) and that fluid enters only a well-established
channel behind the complexity of the process zone.
These aspects of the lag region provide great simplification and increased predictablility for applying
commercial-scale hydraulic fracturing processes.
A paper by Howard and Fast (1957), and particularly the accompanying appendix by R. D. Carter,
provides the current framework for fluid loss. The
paper identifies the three factors controlling fluid
loss: filter-cake accumulation, filtrate resistance into

A5-7

0.6

1.0

Test
4
5
6
7
9
10
11

0.5

0.4

0.9
0.8

0 = 3
16

0.7

0 =
8

w/wo

w/wo

0.6
0.3
Width at
fluid arrival

0 =
4

0.4

0.2

0 =
16

0 = 3
8

0.5

0.3
0.2

0.1

0.1
0
0.25

0
0.20

0.15

0.10

0.05

Normalized distance from tip, (L x)/L

Normalized distance from well, x/L

1.0

1.0

Test
4
5
6
7
9
10

0.8

0.8
0 =
8

0 = 3
8

p/po

0.6

p/po

0.6

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

0.4

0 =
16

0.4
0 =
4

0.2

0.2
0 = 3
16

0
0.5

0
0.4

0.3

0.2

0.1

Normalized distance from tip, (L x)/L

0.2

0.4

0.6

0.8

1.0

Normalized distance from well, x/L

Appendix Figure 4. Comparison of Warpinski (1985) field data (left) and Khristianovich and Zheltov (1955) analysis
(right). wo and po are the wellbore values of width and pressure, respectively; x is the distance from the well.

the reservoir and displacement of the reservoir fluid


(see Fig. 5-17 and Chapters 6 and 8). All three factors are governed by the relation 1/t (where t is
time) for porous flow in one dimension. The coefficient for this relation was termed the fluid-loss coefficient CL. The authors also provided the means to
determine the coefficient for all three factors using
analytical expressions for the filtrate and reservoir
contributions and to conduct and analyze the filtercake experiment, which is now an American Petroleum Institute (API) Recommended Practice.
Also of significance was presentation of the Carter
area equation, with area defined as the product of the

A5-8

height and tip-to-tip length. This equation, based on


the assumption of a spatial and temporal constant
fracture width, provided the first rigorous inclusion
of fluid loss into the fracturing problem (see Chapter
6). Equation 6-18, which is solved by Laplace transformation, is in terms of exponential and complementary error functions and is not engineer friendly.
This difficulty was soon overcome by developing a
table for the more complicated terms in the equation
using a dimensionless variable (see Eq. 6-19) that is
proportional to the fluid-loss coefficient (loss volume) divided by the width (stored volume) and
hence also related directly to the fluid efficiency

Chapter 5 Appendix: Evolution of Hydraulic Fracturing Design and Evaluation

illustrated in Appendix Fig. 2. Nomographs for


the complete equation were also developed (e.g.,
figs. 4-17 and 4-18 of the Howard and Fast Monograph). Eventually a simple and approximate expression (Harrington et al., 1973) for the Carter equation
provided the basis for fracture design into the 1980s.
The approximate expression is based on the relation
at the top of Appendix Fig. 2. For these applications,
the average width was first determined from either the
KGD or PKN model, as discussed in the following.
Another 1957 paper was by Godbey and Hodges
(1958) and provided the following prophetic phrases:
By obtaining the actual pressure on the formation
during a fracture treatment, and if the inherent tectonic stresses are known, it should be possible to
determine the type of fracture induced. . . . The
observation of both the wellhead and bottomhole
pressure during fracturing operations is necessary to
a complete understanding and possible improvement
of this process. These statements anticipated two of
the important enablers for the second generation of
fracturing: the use of pressure in an manner analogous to well test characterization of a reservoir and
employment of a calibration treatment to improve
the subsequent proppant treatment (see Chapters 5,
9 and 10).
In 1961 Perkins and Kern published their paper
on fracture width models, including the long aspect
ratio fracture (length significantly greater than height)
and radial model (tip-to-tip length about equal to
height) as described in Section 6-2.2. They considered,
for the first time, both turbulent fluid flow and nonNewtonian fluids (power law model) and provided
validating experiments for radial geometry and the
role of rock toughness.
Perkins and Kern also discussed fracture afterflow
that affects the final proppant distribution within the
fracture. After pumping stops, the stored compression in the rock acts in the same fashion as compressible fluids in a wellbore after well shut-in. After
fracture shut-in, fluid flow continues toward the tip
until either proppant bridges the tip or fluid loss
reduces the fracture width and stored compression
to the extent that the fracture length begins to recede
toward the wellbore (Nolte, 1991). The magnitude
of the fracture afterflow is large compared with the
wellbore storage case, as discussed later for
Appendix Eq. 4.

Reservoir Stimulation

The one shortcoming acknowledged by Perkins


and Kern was not rigorously accounting for the flow
rate change in the fracture required by continuity
(i.e., material balance). They assumed that the volumetric flow rate was constant along the fractures
length, which does not account for the effects of
fluid loss and local rates of width change (storage
change). This assumption was later addressed by
Nordgren (1972), who provided closed-form equations for the bounding cases of negligible fluid loss
and negligible fracture storage (i.e., most fluid
injected is lost during pumping) for a long-aspect
fracture and Newtonian fluid (see Section 6-2.2). The
initial letters of the last names of the authors of these
two papers form the name of the PKN model.
The remaining paper of historic importance for
width modeling is by Geertsma and de Klerk (1969).
They used the Carter area equation to include fluid
loss within the short-aspect fracture, as previously
considered by Harrison et al. (1954) and Khristianovich and Zheltov (1955). Their initials coupled with
those of the authors of the latter paper form the name
of the KGD (or KZGD) width model.

Reservoir response to a fracture


Until the advent of numerical simulators, production
models for a fracture did not consider transient flow
effects and were based on the FOI relative to the
reservoirs radial flow response with no damage (skin
effect = 0). The increase in production, relative to the
case before fracturing, can be significantly greater
than the FOI measure because fracturing also bypasses near-wellbore damage. The enhanced stimulation benefit increases as the magnitude of the damage
increases. For example, removing a skin effect of
about 25 increases production by about a factor of 4,
whereas during the first generation a typical FOI target
was about 2, relative to zero skin effect.
Papers considering finite-conductive fractures
began to appear in 1958 and are summarized in chapter 10 of the Howard and Fast (1970) Monograph.
Craft et al. (1962) considered the combined effects of
fracture stimulation and damage bypass. Also of historical interest is that most of this work was performed on analog computers with electrical circuits
representing the reservoir and fracture components.
Recognition of the role of conductivity was important
because the idealized assumption of infinite conduc-

A5-9

(J/Jo)

7.13

re

ln 0.472 rw

14
L/re = 1

12
10
8

0.9
0.8
0.7
0.6
0.5
0.4

0.3

0.2
0.1

2
0
102

103

104

Relative conductivity,

105
kfw
k

106

40

Appendix Figure 5. McGuire and Sikora (1960) curves for


folds of increase (J/Jo) in a bounded reservoir of area A
(acres).

A5-10

1.0
Effective wellbore radius, rw/xf

tivity, with no pressure loss in the proppant pack,


cannot result from an economics-based optimized
treatment. The incremental production increase, by
achieving the infinite-acting case, would not offset
the operational cost for the additional proppant.
McGuire and Sikora (1960) presented a significant
study of the production increase in a bounded reservoir for a fracture with a finite conductivity kf w for
the proppant pack, where kf is the fracture permeability. The boundary and conductivity effects are summarized in the set of pseudosteady-state curves
shown in Appendix Fig. 5. The curves reflect different ratios of the fracture length relative to the
drainage radius re, with the vertical axis reflecting
the FOI as J/Jo and the horizontal axis reflecting
dimensionless conductivity based on the drainage
radius. The McGuire and Sikora curves were the primary reservoir tool for fracture design and evaluation until the late 1970s.
Prats (1961) used mathematical analyses to conduct a comprehensive consideration of finite-conductivity fractures with the assumption of steady-state
flow (i.e., constant-pressure boundaries). He introduced a dimensionless conductivity that is essentially the inverse of the dimensionless fracture conductivity commonly used for transient analyses (i.e.,
CfD = kf w/kxf = /2). Prats also introduced the concept of an effective (or apparent) wellbore radius rw.
The effective radius allows describing the fracture
response in terms of an enlarged wellbore radius
within the radial flow equation. This concept is illustrated in Appendix Fig. 6 for pseudoradial flow
(adapted from Cinco-Ley and Samaniego-V., 1981b).

rw = 0.5x f

0.5

1
rw = 0.28

k fw
k

CfD = 30

0. 1

0.01
0.1

CfD =
CfD = 0.2

1.0

k fw
kx f

10

100

Dimensionless fracture conductivity, CfD

Appendix Figure 6. Effective wellbore radius versus


dimensionless fracture conductivity (Nolte and
Economides, 1991, adapted from Cinco-Ley and
Samaniego-V., 1981b).

The effective wellbore radius, coupled with the radial


flow equation, provides a powerful tool for efficiently
calculating the FOI, or negative skin effect, provided
by the fracture. Prats also considered fracture face
damage (or skin effect) and provided an optimized
treatment based on a fixed amount of proppant.

Treatment optimization
Optimizing a fracture treatment is an essential part
of maximizing its benefit (see Chapters 5 and 10).
For this reason Prats (1961) optimization consideration is of historic importance, although proppant volume is generally not a realistic criterion because
proppant cost is only part of the investment for a
fracture treatment (e.g., Veatch, 1986; Meng and
Brown, 1987). Prats proppant optimization condition at CfD = 1.26 could be a practical target for
high-permeability reservoirs; however, this value is
about an order of magnitude lower than the optimum
case for the long transient period of a very low permeability reservoir.
Additional lessons are also provided by the apparentwellbore concept. The first is that a fracture is
equivalent to enlarging the wellbore and not increasing the formations global permeability. Incorrectly
considering a fracture to be a permeability increase
can lead to incorrect conclusions concerning reservoir recovery and waterflood sweep. Another insight
is the generally favorable economics for an effectively designed and executed fracture. A fracture

Chapter 5 Appendix: Evolution of Hydraulic Fracturing Design and Evaluation

treatment is equivalent to excavating a very large


diameter borehole (e.g., hundreds of feet in most
cases) and therefore is an extremely cost-effective
way to provide an equivalent excavation.
The most important optimization lesson is found
in Appendix Fig. 6 for the roles of conductivity kf w
(achieved by proppant cost) and fracture penetration
(achieved by fluid and other additive costs; see
Chapter 7). The figure indicates that as CfD increases
beyond 10, the effective wellbore radius approaches
one-half of the fracture length and there are diminishing returns for additional increases in conductivity
(i.e., incurring proppant costs without an effective
increase in production rate). For this part of Appendix Fig. 6, the effective radius is constrained only
by length and is termed the length-limited case.
However, increasing both fracture length and conductivity to maintain a constant CfD achieves the
most efficient conversion of length into an effective
wellbore radius. This conversion is the basis for
effectively fracturing low-permeability formations.
The practical limits for the length-limited case are
reaching the drainage radius, increasing conductivity
within the limits of achievable fracture width and
efficiently extending a fracture when the pressure
reaches the formation capacity, as discussed later. As
permeability increases, and proportionally decreases
CfD, the ability to increase conductivity becomes
the constraint. As CfD progressively decreases, the
conductivity-limited case is reached. The figure indicates that as CfD decreases below 1, a log-log unit
slope is approached that relates rw to kf w/k, with the
obvious absence of an effect from length. When the
unit slope is reached, near a value of 0.2, the wellbore drawdown completely dissipates within the
fracture before reaching the tip, and the extremities
of the fracture cannot provide a production benefit.
For the conductivity-limited condition, the production rate can be increased economically only by providing more conductivity kf w, with an obvious constraint from the available fracture width developed
during the treatment. This constraint was significantly
extended by the third fracturing generation of TSO
treatments, which is discussed toward the end of this
Appendix.

Reservoir Stimulation

Transition between the first and


second generations
By 1961, the design and evaluation tools for most
of the next two decades had been established by the
contributions discussed. Incremental development of
these tools slowed because fracturing was considered
a mature technology. Also affecting technical development was the degrading economics for lower quality reserves as oil import-export increased and fracturing activity decreased (Appendix Fig. 1). This
condition did not change until the mid-1970s
brought natural gas shortages and higher gas prices
to the United States. Higher prices produced the
incentive to develop extensive regions of tight gas
reserves with fractures targeting the FOI = 10 range
of the McGuire and Sikora curves (Appendix Fig. 5).
Before this period, typical fracturing targets were oil
reservoirs with an FOI of about 2, with FOI relative
to an undamaged wellbore. However the FOI = 10
target required about an order-of-magnitude increase
in the volume and cost for a typical treatment and
was hence termed massive hydraulic fracturing.
This new target introduced higher temperature
reservoirs, typically of tight gas, that generally
exceeded the performance limits for fracturing fluid
systems. These conditions stretched the so-called
mature technology in almost every conceivable way
and resulted in a bumpy journey because of the proportionally large economic penalty when a treatment
failed to meet expectations. However, reports of successful field development (e.g., Fast et al., 1977)
encouraged continued interest in tight gas development.

Realistic estimate of conductivity


Cooke (1975) reported realistic experiments for characterizing the conductivity of proppant packs. His
procedure formed proppant packs from a slurry composed of polymer-based fluids by using a cell with
rock faces that allowed fluid loss and the subsequent
application of closure stress. The Cooke cell is now
a standard apparatus for a fracturing fluid laboratory

A5-11

(see Chapter 8). The experiments showed that the


retained pack permeability could be very small.
These results were unexpected because prior testing
procedures did not use fracturing fluids or stress levels for deeper gas reserves. The primary difference
resulted because the rock acts as a polymer screen
at moderate and smaller permeability levels, which
significantly increases the polymer concentration
remaining within the proppant pack porosity after
fracture closure.
Cooke also provided a simple mass-balance relation for this important consideration. The concentration factor for the polymer and other additives
remaining in the fracture relative to the original
concentration can be expressed as
CF = 44 / ppa ,

(3)

for a typical proppant pack porosity of 0.33 and


proppant specific gravity (s.g.) of 2.65. The relation
depends on the average concentration <ppa> defined
as the total pounds of proppant divided by the total
gallons of polymer-based fluid. This relation indicates
a polymer concentration increase of 20 or greater for
typical treatments at that time (e.g., <ppa> of 1 to
2 lbm). This unexpected discovery of a significant
reduction in retained permeability, coupled with the
prior discussion on conductivity and effective wellbore radius, partly explains the difficult transition to
massive treatments.
Cookes pioneering work had obvious effects on
proppant schedules for treatments and laboratory
testing procedures. Equally important, the work initiated substantial product development activities, as
discussed in Chapter 7. These include improved
proppants, beginning with Cookes work on bauxite
for high crushing stress, improved breaker chemistry
and breaker encapsulation, large reductions of polymer concentration for crosslinked fluids, foams and
emulsions, and residue-free viscoelastic surfactant
systems. The evolution of fracturing fluid chemistry
was reviewed by Jennings (1996).

Height growth and proppant transport


Simonson et al. (1978) presented the mechanics governing fracture growth into a layer with higher stress,
complementing the postulate by Harrison et al. (1954)
concerning the role of stress for height confinement.
The analysis considered a three-layer case for two
symmetric barriers (i.e., two barriers extending

A5-12

infinitely above and below the pay section with each


barrier having the same magnitude of stress). The
three-layer case provided insight into how to adapt
more general relations to any number of layers (e.g.,
Nolte and Smith, 1981; chapter 3 of Gidley et al.,
1989). These relations led to the calculations employed
in pseudo-three-dimensional (P3D) fracture simulators
(see Section 6-3.2).
Novotny (1977) outlined a comprehensive basis for
proppant transport calculations and in particular identified the important roles of channel shear rate and fracture closure in determining the ultimate placement of
proppant (see Section 6-5.3. Both effects produce more
proppant fall. For non-Newtonian fluids, the effective
viscosity for sedimentation is determined from the vectoral sum of the shear rate in the channel and that
caused by proppant fall (as for stagnant fluid). This
sum is generally dominated by the channel flow and
is much greater than that for a particle in stagnant fluid
(i.e., higher shear rate and lower viscosity). In addition,
the closure period prolongs the time for proppant fall
and maintains the channel flow to reduce the effective
viscosity. Novotny also provided a brief analysis of the
volume balance during closure, which is the essential
ingredient for the fracturing pressure decline analysis
(e.g., Nolte, 1979) that is used for calibration treatments (see Section 9-5).

Transient reservoir response


The FOI consideration for fracture production was
found to be completely inadequate for the substantial
period of transient flow that occurs in tight formations
(see Section 12-2). The first tool for finite-conductivity
transient flow was type curves provided by Agarwal
et al. (1979). Although these curves were developed
from numerical simulators, access to computers was
generally outside the reach of most engineers. These
and similar type curves became the standard evaluation tool to assess production from a fracture treatment. Type curves were also used for optimizing treatment design. By the mid-1980s, as general access to
computers increased, the use of type curves began to
decrease accordingly.
Cinco-Ley and Samaniego-V. (1981b) provided
several advancements for understanding and quantifying the transient behavior of a reservoir fracture
system. In addition to advancing the effective wellbore concept (e.g., Appendix Fig. 6) and type curves,
they identified and provided comprehensive descrip-

Chapter 5 Appendix: Evolution of Hydraulic Fracturing Design and Evaluation

tions for the distinctive transient regimes resulting


from a finite-conductivity fracture (see Section 12-2).
The bilinear flow regime, generally the first to
occur during production or a well test, was paramount for bridging the gap between fracture design
and subsequent evaluation based on production or
well tests. For permeability in the range of 10 d, the
bilinear period can last on the order of a year or more
for a long fracture (>2500 ft from the well). During
bilinear flow the stabilized pressure drawdown progresses along the fracture length. During this period,
it is not possible to determine the length of the fracture from a well test or production data because the
total length has not had time to effectively experience
the wellbore drawdown. Therefore, a meaningful
evaluation for fracture length cannot be obtained until
the bilinear period ends and the transient response
progresses toward pseudoradial flow (potentially several years). An obvious implication in this case is that
a standard well test cannot be used to determine fracture length; the length can be determined only from
long-term production data. They also identified
another important aspect of bilinear flow that occurs
because of the transient flow condition within the
proppant pack: the fracture conductivity can be characterized, independent of length and hence most reliably, by the slope of a plot of pressure versus the
quarter-root of time.
Recognition of these consequences for bilinear
flow also explains the difficult transition to the successful application of massive treatments. Well test
interpretations misinformed instead of informed.
They indicated relatively short fracture lengths that
were assumed to be treatment placement failures and
led to the common and contradicting result: how can
1 million lbm of sand be contained in a fracture
length of only 100 ft? Much longer propped lengths
were later substantiated by production data after the
bilinear period had ended (e.g., values of fracture
half-length xf > 5000 ft; Roberts, 1981). Another
contribution to incorrect interpretations was ignoring
Cookes (1975) report of very low retained-pack permeability, which led to overly optimistic estimates of
conductivity and proportionally pessimistic estimates
of length. The coupling of these two factors produced incorrect and negative assessments for many
early attempts to establish massive fracturing as a
viable means of developing tight gas formations.

Reservoir Stimulation

These advancements and insight from Bennett


et al. (1986) for layered formations provide a solid
foundation for the reservoir response to fracturing.

The second generation:


massive fracturing
As indicated in the preceding section, the bumpy
road to successful massive fracturing also included
massive penalties because the cost of a fracture treatment could become equivalent to the well cost. The
combined effect of many companies experiencing
$500,000 treatments that did not provide commercial
wells resulted in a significant investment for fracturing research. One result of this effort is the SPE
Monograph Recent Advances in Hydraulic Fracturing (Gidley et al., 1989). The manuscripts for this
comprehensive volume, with more than 30 contributors, were completed in 1984, only five years after
the 1979 SPE annual meeting provided the first
meaningful number of papers from this research
effort. The papers presented at this meeting were
significant also because they presented a key that
enabled the reliable application of massive fracturing
and rapid progression of the treatment size record
from 2 million lbm in 1979 to more than 7 million
lbm by 1986.
The key was that, for the first time in its 30-year
history, fracturing was considered in a framework
similar to that used for reservoir characterization.
The reservoir framework consists of pressure transient analysis for the flow characteristics, wireline
logs for the formation parameters and geophysics
for the macroview. The 1979 papers include the following (a different reference year indicates the publication date):
Logging: Rosepiler (1979) introduced application
of the long-spaced sonic tool to infer stress in different layers (see prior discussion of stress concerning Appendix Eq. 2 and Chapter 4). Dobkins
(1981) presented improved cased hole logging
procedures for inferring the fracture height that
were also used by Rosepiler to qualitatively validate his novel use of mechanical property logs.
Pressure transient analysis (PTA): Nolte and Smith
(1981) introduced the role of pumping pressures by

A5-13

using a log-log plot as a diagnostic tool (similar to


PTA practice) for fracture growth characteristics,
the role of pressure simulation for quantifying
geometry (including height growth) and the role of
calibrated stress profiles obtained from mechanical
property logs. Nolte (1979) introduced the role of
pressure during the postinjection closing period to
quantify fluid loss and predict fracture width and
length by using a specialized time function in a
manner analogous to the Horner plot. The combination of these two papers provided a foundation
for the common use of the calibration treatment
and pressure-history matching for defining design
parameters (see Chapter 9). Appendix Fig. 7 illustrates the fracturing pressure for three distinct phases: pumping, closing and the after-closure period.
Geophysics: Smith (1979) introduced the role of
mapping fracture trajectories by using surface tiltmeters and borehole passive seismic techniques to
improve reservoir recovery by the correct placement of infill wells (see Section 12-1).
A companion paper in 1980 showed the synergistic benefit when these individual considerations are
unified for tight gas exploitation (Veatch and
Crowell, 1982).

Pressure from bottomhole bomb

Bottomhole pressure, pw (psi)

Inferred pressure

9000

Fracture
treatment

Pressure decline
Fracture
Transient reservoir
closing
pressure near wellbore

8000
7000
6000

Fracture closes on
proppant at well

Net fracture
pressure
pnet = pw pc
Closure pressure
pc = horizontal rock stress

Reservoir pressure

5000
38

40

42

44

46

48

50

56

58

Clock time (hr)

Appendix Figure 7. Bottomhole fracturing pressure


(Nolte, 1982).

Fracturing pressure: analog of reservoir


response
An important component of fracturing pressure
analysis is the closure pressure. The closure pressure
is the datum for the net pressure that constrains the

A5-14

width prediction, provides an analog of the reservoir


pressure and reflects the height-averaged minimum
stress for the pay zone (see Sidebar 9A). The fracture width is proportional to the net pressure. The
data in Appendix Fig. 7, one of the first recordings
of bottomhole pressure during a treatment, are similar to the reservoir response for an injection test with
a pressure increase (pumping) and subsequent falloff
(closing). The injection pressure is governed by the
evolving fracture geometry, and the closure data are
governed by the fluid loss. These two conditions,
respectively, enable characterizing the stored and lost
components of the volume-balance equation shown
in Appendix Fig. 2. After closure, the pressure is
independent of the fracture parameters and depends
on the reservoir response to the fluid lost during the
treatment.
The fundamental analogy between reservoir and
fracturing behavior results because a diffusion-type
process governs both behaviors. The respective reservoir and fracturing equivalents are kh/ w2h/
(transmissibility), where k is the permeability, h is the
reservoir thickness, w is the width, and is the
appropriate fluid viscosity, and ct h/(wE) 1/pnet
(storage capacity of the reservoir), where is the
porosity, ct is the total system compressibility, and E
is the formations elastic modulus. The last expression for storage contains an inverse proportionality to
the net fracture pressure pnet. This can be written in
terms of the fracture volume Vf, fluid pressure pf and
closure pressure pc.
1 dVf
1
1
=
=
Vf dp f pnet p f pc

(4)

1
1 dw
=
for constant h and L.
w dpnet
pnet

(5)

This equation implies that the elastic formation,


compressed to contain the fractures volume, produces a system compressibility analogous to an equal
volume of perfect gas at a pressure equal to the fractures net pressure. The result is a significant storage
capacity considering typical conditions with more
than 1000 bbl for fracture volume and only hundreds
of pounds per square inch for net pressure. The last
storage relation, for constant lateral dimensions, is
important for a TSO, as discussed later.

Chapter 5 Appendix: Evolution of Hydraulic Fracturing Design and Evaluation

Fracture simulators
Describing a hydraulic fracture produces a significantly more complex role for the diffusive process
than the reservoir case because the basic parameter
groups change continuously with time, with a nonlinearity for the equivalent permeability, and the farfield elastic coupling between width and pressure
produces local parameters that have a general dependence on the pressure everywhere within the fractures unknown boundaries. For these reasons, fracture simulators that rigorously and robustly couple
these parameters in a general manner (see Section
6-3) have not progressed at the same rate as reservoir
simulators.
The modeling difficulties led to widespread use of
simulators based on P3D assumptions that partially
circumvent the far-field elastic-coupling condition.
The two most common means were relaxing the lateral coupling in the long direction of the fracture (as
for the PKN model) to allow a cellular representation
and vertical height growth of the cells (e.g., Nolte,
1982) or prescribing the boundary and width profiles
by elliptical segments and a lumped dependence on
the governing parameters (e.g., Settari and Cleary,
1986). P3D models, or more precisely P2D models,
evolved to include automated proppant scheduling
and the temperature-exposure history for polymer and
additive scheduling (e.g., Nolte, 1982), acid fracturing (e.g., Mack and Elbel, 1994), economic optimization for treatment design (e.g., Veatch 1986; Meng
and Brown, 1987), automated pressure-history matching (e.g., Gulrajani and Romero, 1996; Gulrajani et
al., 1997b) and rigorous 2D slurry flow (e.g., Smith
and Klein, 1995).
Originally restricted to in-office use, these models
merged with on-site fracture monitoring systems to
provide treatment evaluation and simulation in realtime mode. An equally important advance was the
parallel evolution of process-controlled mixing and
blending equipment for reliable execution of more
demanding treatment schedules and progressively
more complex chemistry that requires precise proportioning (see Chapters 7 and 11).

Fracture mapping and model validation


An important achievement was the definition of fracture length, height and width by employing passive
seismic measurements and tiltmeters in observation

Reservoir Stimulation

wells (Warpinski et al., 1996; see Section 12-1). The


importance of these measurements for fracture design
and evaluation cannot be overemphasized. Independent measurements for each component of the fracture volume (Appendix Fig. 2) provide a long-awaited
benchmark for validating fracture models.
Like the first generations failure to find a consensus for width models (e.g., Perkins, 1973), pressurehistory matching could not resolve the second generations conflicting adaptations of the P3D framework
(see Chapter 6). The convergence of modeling
assumptions failed for several reasons. The first was
fundamental to the pressure-matching process and
results because of the multitude of opportunities for
nonuniqueness. Another reason was the failure to
achieve a dominant industry opinion on either the
technique or procedures for a specific technique to
define closure pressure (e.g., Plahn et al., 1997). This
state of affairs allowed selecting a closure pressure
procedure to validate particular modeling assumptions and therefore justify relatively arbitrary and ad
hoc modeling assumptions. Techniques to determine
the closure pressure are discussed in Section 3-6 and
the Appendix to Chapter 9.
Because of nonuniqueness in the reservoir response
and the basing of reservoir models on overly idealized
modeling assumptions for a fracture, the reservoir
response cannot generally provide an effective
constraint on the achieved fracture length (Elbel and
Ayoub, 1991; Nolte and Economides, 1991). Mapping
constraints on all three fracture dimensions provide a
unique, objective test of the geometry model assumptions (e.g., Gulrajani et al., 1997a) and a basis for
rationally judging and selecting the model complexity
appropriate for the specific application, available data
and simulation resources.

Treatment design and evaluation


The primary fracture evaluation advance from the
massive treatment generation is the calibration treatment performed before the proppant treatment to
define placement parameters. Combining the calibration treatment and the purpose-designed TSO treatment produced the primary treatment innovation of
the second generation. The calibrated TSO treatment, developed by Smith et al. (1984), became the
key to the third fracturing generation (discussed
later) and essentially removed width as a constraint
for the conductivity required to successfully fracture

A5-15

The following paragraphs link several aspects of the


massive and TSO generations by using the information available from the diagnostic log-log plot for
fracturing in Appendix Fig. 8. Appendix Table 1 lists
the interpretations for various slopes exhibited in the
figure by the net pressure during fracturing. The data
are from two massive treatments in tight gas formations. The top curve is a treatment in the Wattenberg
field, the first microdarcy-permeability field development (Fast et al., 1977). The behavior shown by the
lower treatment curve, which was designed by this
author, provided insight for developing the TSO
treatment that enables successfully fracturing darcyscale oil formations. The treatment related to the
lower curve was not particularly successful. However, it was one of the first 2 million lbm treatments
and hence functioned better as a sand-disposal
treatment than a gas-stimulation treatment. The sand
was disposed of with 900,000 gal of crosslinked fluid
containing 90 lbm/1000 gal of polymer, or approximately 80,000 lbm of polymer.
The marginal success of the treatment is readily
understood by considering Appendix Eq. 3. For the
treatment average of 2.1 ppa, the equation predicts
1900 lbm/1000 gal crosslinked fluid (in reality, a
solid) remaining in the proppant pack porosity after
the treatment. However, the size and viscosity for
this treatment provided an ideal test condition of
how a formation responds to fluid pressure and an
excellent illustration for the concept of formation

Variable injection rate

2000
Proppant
begins

II

1000

II

500
40

III-a

Proppant
begins

II

[5 MPa]

60

III-b

IV
III-a

100

200

400

600

1000

Time (min)

Idealized Data
log pnet

Transition between the second and


third generations

Field Data
Net pressure, pnet (psi)

very high permeability formations. This capability


and timing produced the overly optimistic prediction
in 1985 for the beginning of the TSO generation, as
indicated by Appendix Fig. 1a.

III-b
II
III-a

IV
Inefficient extension for
pnet formation capacity pfc

log time or volume

Appendix Figure 8. Log-log diagnostic plot for fracturing


(Nolte, 1982).

capacity. The capacity (Nolte, 1982) defines the pressure limit for efficient fracture extension and is analogous to the pressure-capacity rating for a pressure
vessel. The cited reference has an unsurprising
theme of the negative effects of excesses of pressure,
polymer and viscosity.
Three mechanisms for a formation can define its
pressure capacity before rupture accelerates fluid
loss from the formations pay zone. The subsequent
fluid loss also leaves proppant behind to further
enhance slurry dehydration and proppant bridging.
Each mechanism is defined by the in-situ stress state
and results in a constant injection pressure condition,
or zero log-log slope, when the net pressure reaches
the mechanisms initiation pressure. The mechanisims are

Appendix Table 1. Slopes of fracturing pressures and their interpretation in Appendix Fig. 8.
Type

Approximate log-log slope value

Interpretation

8 to 14

Restricted height and unrestricted expansion

II

Height growth through pinch point, fissure opening


or T-shaped fracture

III-a

Restricted tip extension (two active wings)

III-b

Restricted extension (one active wing)

IV

Negative

Unrestricted height growth

A5-16

Chapter 5 Appendix: Evolution of Hydraulic Fracturing Design and Evaluation

opening the natural fissures in the formation, governed by the difference in the horizontal stresses
extending the height through a vertical stress barrier and into a lower stress (and most likely permeable) zone, governed by the difference in the
horizontal stress for the barrier and pay zone
initiating a horizontal fracture component when
the pressure increases to exceed the level of the
overburden stress.
An important observation for the pressure capacity
is that it depends on the in-situ stress state and therefore does not change for the formation in other well
locations unless there are significant local tectonic
effects. As a result, all future treatments for the field
can generally be effectively designed on the basis of
only one bottomhole pressure recording and its
detailed analysis (see Section 9-4).
The upper curve on Appendix Fig. 8, for the
Wattenberg treatment, illustrates the fissure-opening
mechanism with the Type II zero slope occurring at
a net pressure of 1700 psi. This value provides one of
the largest formation capacities ever reported. The fissure opening is preceded by restricted height growth
and unrestricted extension (Type I slope) that provide
the most efficient mode of fracture extension. Therefore, conditions in this formation are favorable for
propagating a massive fracture; not by coincidence,
this was the first field successfully developed in the
massive treatment generation (Fast et al., 1997), and it
provided incentive to continue the development of
massive treatment technology. Returning to Appendix
Fig. 8, after the period of constant pressure and
enhanced fluid loss, a Type III-a slope for a fracture
screenout occurs because slurry dehydration forms
frictional proppant bridges that stop additional extension (i.e., a generally undesired screenout for a tight
formation requiring fracture length over conductivity).
After the penetration is arrested, the major portion of
the fluid injected is stored by increasing width (see
Appendix Eq. 4) and the net pressure develops the unit
slope characteristic of storage. The amount of width
increase is proportional to the net pressure increase.
The Wattenberg treatment consisted of 300,000 gal
of fluid and 600,000 lbm of sand with an average
concentration of 2 ppa, similar to the previous example. However, the treatment was successful because
a polymer-emulsion fluid with low proppant pack
damage was used. After the treatment defined the
formation capacity, model simulations indicated that

Reservoir Stimulation

the required penetration could be obtained by not


exceeding the formation capacity. A subsequent treatment designed using 150,000 gal and 900,000 lbm of
sand (an average of 6 ppa) became the prototype for
the remaining development of the field (Nolte, 1982).
The lower curve on Appendix Fig. 8 is for the
aforementioned sand-disposal treatment in the Cotton Valley formation of East Texas. As previously
discussed, the treatment provided an opportunity to
observe a large range of fracturing behavior with five
types of interpretive slopes occurring, including
Type I indicating extension with restricted height
growth
Type II defining this formations lowest pressure
capacity at 1000 psi for the penetration of a stress
barrier
Type IV, with decreasing pressure, indicating unrestricted vertical growth through a lower stress
zone after the barrier was penetrated.
The Type IV condition continued until proppant
was introduced. Almost immediately after proppant
entered the fracture the pressure increased, most
likely because the proppant bridged vertically in the
width pinch point formed by the penetrated stress
barrier and restricted additional height growth.
During the preceding 6-hr period of significant vertical growth, the horizontal growth was retarded. As a
result, the very high polymer concentration formed a
thick polymer filter cake at the fracture tip that probably restricted further horizontal extension. Thus, the
extremities of the fracture were restricted either by
proppant or polymer cake, and continued injection
was stored by increasing width indicated by the Type
III-a unit slope. After a significant increase in pressure, the pressure became constant for a short period
at 1200 psi with a Type II slope that probably resulted
from opening natural fissures to define a second,
higher capacity. Subsequently the slope increased to
an approximately 2:1 slope indicated as Type III-b.
This latter slope for a storage mechanism indicates
that about one-half of the fracture area had become
restricted to flow, which could have resulted from one
wing of the fracture being blocked to flow near the
well because of slurry dehydration from the fissure
fluid loss. The wellbore region experiences the largest
pressure and is most prone to adverse fluid-loss
effects from exceeding a capacity limit.

A5-17

Subsequent treatments were improved after understanding the formations pressure behavior as in the
Wattenberg case and for this area after understanding
the implications of Appendix Eq. 3 for concentrating
polymer. In addition, the observation that proppant
bridging could restrict height growth was developed
for treatments to mitigate height growth (Nolte,
1982). An effective and relatively impermeable
bridge can be formed within the pinch point to retard
height growth by mixing a range of coarse and fine
sand for the first sand stage after the pad fluid.
Smith et al. (1984) later sought a means to significantly increase fracture width for the development of
a chalk formation within the Valhall field in the
Norwegian sector of the North Sea. The additional
width was required because laboratory tests indicated
the likelihood of substantial proppant embedment
into the soft formation that would lead to the loss of
effective propped width. Fracturing was considered
for this formation because other completion techniques would not sustain production because of chalk
flow. The resulting treatment design was based on the
behavior on the log-log plot in Appendix Fig. 8 for
the sand-disposal treatment: a purpose-designed TSO
treatment. For the disposal treatment, they observed
that after the initial screenout occurred, 2 million lbm
of proppant could be placed, and the net pressure
increase indicated that this occurred by doubling the
width after the screenout initiated.
Smith et al. designed and successfully placed a
TSO treatment in which proppant reached the tip and
bridged to increase the width by a factor of 2 during
continued slurry injection after the purpose-designed
TSO occurred. This design, with successful placement of progressively larger propped width increases,
became the tool that enabled the development of this
formation. The ability to significantly increase the
width after screenout results from the large storage
capacity of a fracture, as detailed in the discussion
following Appendix Eqs. 4 and 5. Additional discussion on the fracture completion in Valhall field and
the TSO treatment is in the Reservoir and Water
Management by Indirect Fracturing section.
As a historical note, a similar concept for a TSO
was disclosed in a 1970 patent (Graham et al.,
1972), with the bridging material consisting of petroleum coke particles (approximately neutral density to
ensure transport to the extremities). The patents goal
was increased width to enable placing larger size
proppant in the fracture.

A5-18

The third generation: tip-screenout


treatments
A proper historical perspective of this third generation requires perspective from the next generations;
however, several of its developments are reviewed
here. A more comprehensive presentation and reference are by Smith and Hannah (1996).
Demonstration of the ability to routinely place a
successful TSO treatment opened the door for effective fracture stimulation of higher permeability
formations. Another component for the successful
fracturing of high permeability was the continued
development of synthetic proppants that can produce
a cost-effective 10-fold increase in permeability relative to sand for higher closure stresses (see Chapter
7). Coupling this increase in permeability with the
similar increase for propped width achieved by a
TSO treatment in a moderate- to low-modulus formation provides about a 100-fold increase in conductivity over a conventional sand fracture. The
conductivity increase also translates into a 100-fold
increase of the target permeability for fracturing, as
implied by Appendix Figs. 5 and 6. The increases for
width and conductivity also mitigate nondarcy (or
turbulent) flow effects in the fracture for high-rate
wells, particularly gas wells (see Sections 10-7.3 and
12-3.1).
However, the anticipated growth rate shown on
Appendix Fig. 1a was slowed not only by the unanticipated, extensive contraction of activity in general,
but also by two prevailing mind sets: high-permeability formations cannot be successfully fracture
stimulated and why fracture a commercial well?
Additional field proof for the benefits of a TSO treatment came from two successful programs: a significant improvement over conventional fracture treatments for the Ravenspurn gas field in the southern
North Sea (Martins et al., 1992b) and high-permeability applications in the Prudhoe Bay field (Hannah
and Walker, 1985; Reimers and Clausen, 1991;
Martins et al., 1992a).

Deep damage
Fracturing in Prudhoe Bay was particularly successful
because deep formation damage induced by prior production (i.e., beyond the reach of matrix treatments)
facilitated sidestepping the mind set of not applying

Chapter 5 Appendix: Evolution of Hydraulic Fracturing Design and Evaluation

fracturing to high permeability. The incremental production from only one year of the fracturing program
would have ranked as the 10th largest producing field
in the United States (e.g., Smith and Hannah, 1996),
without including similar results achieved by another
operator in the other half of the field. Another significant aspect of the Prudhoe Bay application is that the
fractures were routinely placed in a relatively small oil
zone above a rising water zone without entering the
water zone (Martins et al., 1992a), which demonstrated that fracturing is a viable, potentially superior
alternative to matrix treatments in high-permeability
formations. This precise fracturing was achieved by
coupling an initial detailed fracture modeling study
with a calibration treatment before each proppant
treatment.

Frac and pack


The frac and pack completion consists of a TSO
treatment before a conventional gravel pack. During
the early 1990s, frac and pack treatments were
applied on a limited basis around the world, notably
offshore Indonesia. Prior to the TSO treatment era,
this technique was tried at various times but without
sustained success. The large propped width from a
TSO treatment was a necessary ingredient for successful frac and pack applications, as discussed later.
The frac and pack boom was in the Gulf of
Mexico. The first successful application began
because of economic considerations and therefore
overcame the mind set of not fracturing a commercial well. A significant field development was not
meeting production expectations because standard
gravel-packed completions could not consistently
achieve a low skin effect; the skin effect ranged
between 7 and 30. The skin effect was 10 after the
first frac and pack treatment and progressively
decreased to near zero from improvements in the
treatment design and the use of larger size proppant
(Hannah et al., 1994).
The threefold-plus increase in production rate, by
eliminating the skin effect, resulted from more than
just adding a TSO treatment to the procedure. An
important feature of a frac and pack is reduction of
the inherent flow restriction around and through the
perforations. The ring of proppant around the casing
(Appendix Fig. 9) acts as an excellent external gravel
pack for reducing the pressure drop through the perforated region. The ring results from the large TSO

Reservoir Stimulation

fracture width that mechanically must continue


around the wellbore; i.e., if the formation is pushed
apart 2 in. over the large surface area of the fracture,
the rock around the wellbore must be displaced
accordingly. For a well-designed and executed frac
and pack, the initiating screenout at the tip is progressively packed back to the well to completely
pack the resulting ring.
The continuing success of the initial frac and
packs started a rapid conversion to this completion,
with the frac and pack becoming the preferred Gulf
of Mexico sand control completion. In addition to
continued use offshore Indonesia, technology transfer resulted in a wider geographical distribution for
this sand control technique (e.g., West Africa,
Gulrajani et al., 1997b).
As for other applications of TSO treatments, on-site
redesign after a calibration treatment became a standard frac and pack practice. An important observation
is that the same analysis procedures and design models introduced for the massive treatments of tight gas
formations in the late 1970s were transferred directly
to frac and pack treatments in soft formations.

Casing

External gravel pack


connecting all perforations
with propped fracture

Packed-back
fracture

Appendix Figure 9. Successfully packed-back TSO treatment.

Reservoir and water management


by indirect fracturing
Another application of TSO treatments is reservoir
management. The prototype example for this application was in the Norwegian Gullfaks field (Bale et al.,
1994a, 1994b). The reservoir section had a multidarcy-permeability upper zone that graded downward
to a permeability of about 100 md. The standard completion was to perforate and gravel pack the upper
zone. However, an edge-water drive would encroach
through the high-permeability zone and turn a prolific
oil well into an even higher water producer.

A5-19

A solution was found from the pioneering work


of the Valhall TSO treatment discussed for Appendix
Fig. 8. This application in the early 1980s was for
more than mitigating proppant embedment. The primary objective was for controlling chalk production
from the primary producing zone above where the
TSO treatment was placed. The upper chalk zone
was very soft with high porosity and composed of
almost as much oil as chalk. When this zone was put
on production, chalk filled the tubing and led to casing collapse. The zone was produced by placing the
TSO treatment in the more competent zone below
and extending the fracture height into the bottom of
the very high porosity formation. This completion
enabled chalk-free production from both the upper
and lower zones (Smith et al., 1984).
This indirect access to the primary producing zone
has come to be known as an indirect vertical fracture
completion (IVFC) and is illustrated in Appendix
Fig. 10. The technique of perforating and fracturing
only from competent sections and producing from
incompetent sections is a robust method for controlling the production of formation material and
increasing recovery from the lower permeability
zones by fracture stimulation. From this perspective,
a TSO-IVFC becomes a solids control and reservoir
management application (see Section 5-1.2).
The Gullfaks adaptation by Bale et al. (1994a)
also placed a TSO-IVFC in a lower competent part
of the formation. In addition to providing sand control and managing reservoir depletion, it was a water
management treatment because it delayed water
breakthrough and greatly increased reserves recovery

High
permeability

Propped
fracture

Low or
moderate
permeability

Appendix Figure 10. Indirect vertical fracture for reservoir


management (Bale et al., 1994a).

A5-20

from the lower sections by fracture stimulation and


a significant increase in drawdown. This application
completes the link between the sand-disposal thight
gas treatment in Appendix Fig. 8 to reservoir and
water management with the intermediate development of the TSO-IVFC for solids control in the
Valhall field.

Screenless sand control


Another apparent role of the IVFC is to eliminate the
need for a screen in many sand-control environments
by selecting and perforating only competent sections
within or near the unconsolidated sections of the formation. The zone selection method can potentially be
enhanced by a sonic log application. This application
takes advantage of the generally considered negative
effect of near-wellbore refracted and relatively slower
waves caused by the wellbore mechanical damage
that routinely occurs in weak or highly stressed formations (Hornby, 1993). However, for screenless
completions, the negative effect becomes a positive
effect because the change in the wave speed for the
refracted wave is a direct indication of the state of
rock failure around the well, which is caused by the
wellbore stress concentration within the in-situ stress
field. Therefore, the layers with a minimal near-well
change in wave speed relative to the far-field speed
are the more competent candidate zones for perforating and applying a TSO-IVFC to achieve screenless
formation-material-controlled production.
A second method of achieving a screenless sandcontrol completion is applied without strategically
placed perforations (e.g., Malone et al., 1997). This
method couples the proppant ring around the casing
from a TSO treatment and proppant with effective
flowback control (e.g., fibers, curable-resin-coated
proppant or both). The combination with a successful packed-back TSO achieves an external gravel
pack of stable proppant (i.e., an external formationmaterial screen as illustrated by Appendix Fig. 9).
Perforation and completion considerations are
addressed in Section 11-3.5.
The screenless completion obviously eliminates
the cost of the screen and related tools, but more
importantly it enables economic development of significant behind-pipe reserves that do not warrant the
mobilization and operational costs for a rig on an
offshore production platform, as generally required
for a standard gravel-pack completion.

Chapter 5 Appendix: Evolution of Hydraulic Fracturing Design and Evaluation

A future generation: fracturing and


reservoir engineering merger?
The previous discussion of the TSO generation clearly
shows the blurring of what can be controlled on the
inside and outside of the casing and of what have
been the traditional roles of a fracture design engineer
and a reservoir engineer. This blurring of past distinctions provides prospects for additional innovations
and the advent of a fourth fracturing generation.

Optimal reservoir plumbing


From a broader viewpoint, the IVFC and strategically
placed perforations provide the means to extend optimized plumbing into the reservoir. Optimized plumbing, through a NODAL analysis, is generally practiced
only for the surface facilities and within the wellbore.
Extended optimization requires additional considerations for designing the plumbing system provided by
the fracture in the reservoir and also within the fracture itself.
The outline for these considerations was defined by
Bale et al. (1994b) for the Gullfaks application. They
considered the role of the permeable fracture plane on
the reservoirs 3D flow pattern and how tailoring the
distribution of conductivity can advantageously affect
this flow pattern (e.g., reducing the conductivity as the
fracture approaches the high-permeability upper zone
to delay water production while increasing the conductivity in the lower permeability zone and applying
a large drawdown to accelerate production from this
zone; see Section 5-1.2). Therefore, the analysis and
design tools have evolved for considering the role of
fractures in NODAL analysis for reservoir, formation
material and water management.

Achieving full potential for horizontal


wells and laterals
The preceding discussion of the IVFC is in the context of single, essentially vertical wells. The potential
for innovative strategies to drain a reservoir increases
several fold by adding consideration of horizontal and
lateral wells. These highly deviated wellbores are typically placed without cemented casing because of economic considerations and therefore do not generally
reach their full potential because they lack an effective
technique to remove wellbore damage. The solution

Reservoir Stimulation

of using cemented casing for effective treatment


diversion tends to be overlooked because of an apparent failure to appreciate lifecycle economics or the
effectiveness of good cementing techniques (see
Chapter 11). Staged fracturing, from correctly placed
perforated sections, enables highly effective damage
bypass, as demonstrated by the first fracturing generations rate of 100 treatments per day in 1955.
The general benefit for a horizontal well, particularly
with vertical variations of permeability, is magnified
by the fracture adding a large vertical permeability
component (see Chapters 11 and 12). Simply stated,
an extended reach well cannot drain what it is not connected to nor can it efficiently drain what it is isolated
from by wellbore damage. The addition of a vertical
fracture allows efficient drainage of all isolated sections that the propped fracture reaches. The location
of the fracture, or plumbing source, can be specified by
correctly placed perforations within a cemented casing
and an effective fracture design and execution. Cased
hole logging and logging while drilling can be used
to identify IVFC target locations for connection to bypassed reserves and management for their exploitation.

Fracturing for well testing


The after-closure portion of Appendix Fig. 7, labeled
transient reservoir pressure near wellbore, shows the
return of the fracturing pressure to the reservoir pressure and demonstrates the well testing potential for
any injection above fracturing pressure. This potential
for a fracture is ensured by the well-known result that
the long-term reservoir response is pseudoradial flow
(e.g., Cinco-Ley and Samaniego-V., 1981b) and is the
same flow regime used for standard well testing. An
attractive aspect of the use of fracturing for testing is
that the fracture enhances the likelihood that all the
zones are open and captured by the test. This is an
important consideration for layered formations and
particularly thinly layered zones that can be missed
by open perforations. Another attraction of fracturing
or injection testing is that the wellbore is generally
filled with water that provides minimal wellbore storage and formation volume factor effects.
The long-term radial response following fracture
closure was developed and presented in a pair of
papers: Gu et al. (1993) from the application perspective and Abousleiman et al. (1994) from the theoretical perspective. They recognized that the radial

A5-21

response from fracture injection met the assumptions


for a slug (or equivalently an impulse) test and that
they could directly apply this developed area of reservoir technology.
Another well-known flow regime for a fracture is
pseudolinear flow. Incorporating the analysis of this
after-closure flow regime was the last link of the fracturingpressure analysis chain between the beginning
of injection and returning to reservoir pressure.
Consideration of this regime by Nolte et al. (1997)
indicated that the reservoir memory of the fracturing event can validate several aspects for analysis of
a calibration treatment (e.g., closure time and hence
the critical closure pressure, fracture length and hence
the fluid-loss coefficient, and the division of fluid loss
between normal wall diffusion and tip spurt). Quan-

A5-22

tifying spurt loss is particularly important for highpermeability formations and is not practically attainable by any other means than after-closure analysis.
The after-closure analyses are presented in Section
9-6, and a method to quantify reservoir parameters
during the closure period is presented in Section 2-8.
These applications from the reservoir behavior of
fracturing complement the 1979 adoption of reservoir
methodologies and achieve a direct merging of fracturing into the classic realm of reservoir testing and
characterization (see Chapters 2 and 12). Reservoir
characterization from a calibration testing sequence
to define fracturing parameters provides the ingredients essential for on-site, economics-based treatment
optimization.

Chapter 5 Appendix: Evolution of Hydraulic Fracturing Design and Evaluation

Das könnte Ihnen auch gefallen