Sei sulla pagina 1di 12

Fracture Geometry

Introduction
This section reviews the basic physics of hydraulic fracture propagation, growth, and final fracture
geometry. In addition to leading toward a general understanding of fracture growth, reviewing the basic
physics emphasizes which variables are critical to treatment design. As with all real world processes, the
behavior of hydraulic fracture is a complex function of many variables, however, for the most processes,
a minor number of these variables dominate the process and control; the results. These are critical
parameters or variables, and good information must be available for these critical variables in order to
understand and design the process (in this case to design a hydraulic fracture treatment). For fracturing,
the critical parameters are height, (or the in-situ stress which controls the height); Youngs modulus, ;
fluid loss coefficient, ;and (sometimes) fracture toughness, , e.g.; HECK. Other variables which are
generally less important in controlling fracture geometry include pump rate, , and fluid viscosity,.
Finally, it should be noted that fundamental fracture geometry models discussed in this section from
theoretical basis for Fracturing Pressure Analysis.
Continuity Equation

Determining (or designing) a created hydraulic fracture length can be subdivided into two brad
categories- fracture geometry and fluid loss. On a volume basis this is written as

= + (1)
Which, rearranged, states that the end product of fracture volume is equal to the volume pumped minus
the volume which leaks-off to the formation during pumping. This is the volume balance or material
balance equation, or continuity equation, and the components can be broken down as seen in Eq. (2)

= =
(2)

3 + 2
(3)



. . , (4)
end product,
Combining these gives an approximate relation for one of the major design variables, fracture length.

1 PETE 4090-Unconventional Gas Reservoirs



= (5)
3 + 2 +

where is pump rate in ft3/min (5.614 BBL/min) is total pump time in minutes. C is the fluid loss
coefficient in units of feet/min (with typical values of .0005 to .1). Hp
is the permeable or leak-off height of fracture while is the total gross height of fracture in feet, Sp is
spurt loss in ft3/ft2(.00134 gal/100ft2), with typical values of 0 to 20 gal/100ft2, and w is the average
fracture width (averaged over fracture length and over fracture height ) in feet.
This simple relation is at the heart of all fracture design and is, in itself, a powerful tool. By assuming a
nominal fracture width (say inch = .02 feet 0.6 cm, this simple equation can give surprisingly good
estimates of how altering input data will affect fracture length.
Eq. (5) also illustrates the principal variables in fracture design: 1), width, which is discussed below;2)
C Hp leak off capacity or fluid loss coefficient times leak-off time;3) H, gross or total fracture height.
If H=Hp, as is often the case, then generally height becomes the dominant variable governing fracture
length as seen in Fig. (1).The effect of leak off or fluid loss coefficient, C on fracture length is illustrated
in Fig (2). Note that fracture length versus height and length versus C , in the figures are based on two
dimensional fracture geometry models, which are discussed in more detail below.

Fracture Area
In the previous section fracture geometry, or fracture area, was identified with three terms , and .
For the particular case of a fixed height fracture (and ignoring spurt loss for the moment), the area would
be equal.
= (6)
and the fluid loss area would equal
= = = (7)

2 PETE 4090-Unconventional Gas Reservoirs



where is the ratio of the fluid loss area of the fracture to the total fracture cross-sectional area ( =
or

more generally =
).

The first fracture area (or fracture growth) equation was from Carter [1] and this implicitly inclined
= 1.

2 2
= 2 () + 1,
4
(8)

=2


Harrington [2], et al, in 1973 developed a general relation for fluid volume loss (dale to "matrix" or
"solid loss coefficient" loss) as
= 8 (9)

and Nolte [3] showed that under various bonds the coefficient ( 8 in this case) of this equation could
only vary between 8/3 and 8. Thus, for = 1, and no spurt loss,
3 (10)
re-including spurt loss revises this equation to
3 + 2 (11)
and using the material balance equation gives

= (12)
3 + 2 +

and this is an excellent approximation to the Carter area equation.

Fracture Geometry
The above discussion arid the area equations leads to an extremely important concept in fracturing, the
fracture fluid efficiency. This is simply the fracture volume divided by the total volume pumped. Using
the equations above this can be written as


=
==
3 + 2 +
(13)
=
which is a powerful number for treatment scheduling (e.g., pad volume determination and proppant
addition scheduled. It is also interesting to note how fluid efficiency varies with time, and how "spurt
loss" effects florid efficiency. Also, note that fluid efficiency is referred to as a single number (since
efficiency varies with time during pumping), the reference is to the efficiency just at shutdown.

3 PETE 4090-Unconventional Gas Reservoirs


Assume a nominal fracture width (increasing with time) and a medium value for fluid loss

coefficient, , of .002 . Using these valves, fluid efficiency would vary with time as seen in Table
min

(1) below for two values of spurt loss. (0 to 20 )
100 2

Reviewing the equations above, it is clear that for spurt loss equal to "0", then for small values of time the
fluid efficiency will always be quite high (e.g., for a time of "0", efficiency will be 100%) and that
efficiency will be monotonically decrease with time. However, for a "non-0", value of spurt, the early
time efficiency will be lower (possibly much lower), and during early stage of fracture growth, fluid
efficiency increases as fracture width increases. As time increases the fluid loss coefficient, C, tern will
become more dominant, and again (even with spurt loss) fluid efficiency will begin to decrease. Just as
with many aspects of fracturing, behavior can vary from place-to-place depending on the exact formation
variables!

Fracture Width
If a Slit inside an elastic rock formation is opened by internal pressure, then that silt will open up into
an elliptical cross section crack (Fig. (3)) with a maximum width given by [4]
2 ( )
= =2 , (14)


where = . With having typical values (for rocks) of .15 to .25, is essentially equal to for all
12
practical purposes.
This fundamental elasticity solution was applied to hydraulic fracturing, but it was applied in two
different ways. This resulted in two competing fracture width models, and there is still debate over
which is correct. The question of which model to use is very important since they can predict quite
different fluid, pad, and proppant volumes needed to achieve a required stimulation as seen in Table (2).
NOTE: This example is for a low efficiency (fracture volume/injected volume) case. For low efficiency,
length is dominated by fluid loss and is thus almost width model independent. For higher efficiency, the
resulting lengths of different models would differ much more significantly.
The two width models are commonly called the Perkins & Kern [5] PKN or Geertsma-deKlerk [6] GdK
models and are illustrated in Fig. (4) (along with a third "width model" for radial or penny shaped
fractures for horizontal fractures or vertical fractures with no height confinements). Some basic
mathematical development for a Geertsma-deKlerk, GdK fracture model was proposed by a Russian,
Khristianovic [7]; thus often this fracture geometry is referred to as the KGD model (Khristianovic-
Geertsma-deKlerk).
Though there is debate, general evidence is that the PKN model is more applicable. In a 1976 article [8]
Geertsma concluded For practical application these basic differences suggest that the Perkins and Kern
PKN model is most appropriate for length/height ratios much in excess of unity, while the GdK model is
most appropriate for small values of L/H. "This analytical conclusion was later supported by fracture
width measurements made with a down hole television camera [9], and by mine back experiments done
by Sandia Laboratories [10]. This is also seen in a comparison of calculated fracture widths in Fig. (5).

4 PETE 4090-Unconventional Gas Reservoirs


5 PETE 4090-Unconventional Gas Reservoirs
This figure compares fracture width as calculated for a 3-dimensional fracture with the width would be
calculated by each of the two 2-dimensional theories. Net pressure was held constant for all the
calculations.
The GDK model essentially assumes a fracture is free to slip at the top and the bottom. This is usually
an unreasonable assumption for fracture with L/H (tip to tip length/height) >1. The PKN model, on the
other hand ignores some important fracture propagation/fracture-mechanics principles at the fracture tip.
However, in spite of this, the available data indicates that the PKN model is generally more applicable
where L>H. For L = H, a radial fracture model would be appropriate. For L<H a GdK model would be
correct. However with typical perforated intervals of maybe 50 to 100 feet ( 15 to 35 m) oil field
hydraulic fractures, meeting the criteria of L < H are generally rare.

The correctness of ignoring the fracture tip considerations for PKN fracture geometry (e.g. L > H) in most
situations arises from the fact that net pressure inside a hydraulic fracture (and thus fracture width) is a
result of interaction of fluid flow in the fracture, and the elasticity of the rock. In effect, viscous properties
of the fluid flowing down the entire length of a fracture are generally more dominant than rock
mechanics considerations near the fracture tip. For a PKN model, this relation between fluid flow/fracture
width is discussed below. The relation between fluid flow and fracture width is summarized for all the
remodels in Table (4). [NOTE - Ignoring detailed fracture mechanics processes near the fracture tip is all
right for many cases, particular for fractures with some height confinement and for fractures in normal
rock. However, for other cases, fracture toughness and fracture tip effects can dominate the process. This
is particularly true for radial fractures and fractures in very soft (e.g. low modulus) rocks, and this is
discussed in more detail later.
For a Newtonian viscous fluid, in laminar flow down narrow silt, pressure drop is given by
( ) 12
= = (15)
3

For the PKN model, the elliptical cross section is assumed to be a vertical cross section of the fracture
(e.g., the fracture is closed at to top and bottom) so that
2

= = = (16)
4 4
The fluid flow and width equations can be parameterized as

12
3 = (17)

and
2

= (18)

And these combined to give
1
4 (19)


6 PETE 4090-Unconventional Gas Reservoirs


(Note: for a non-Newtonian power law fluid this can be written as

1
2 +2
(19)

( 1)

where is the consistency index, and is the non Newtonian flow index for a power law fluid. For a
Newtonian fluid = 1, = and this reduces to Eq.
This can be recombined with the width equation to predict the net pressure
3
4 (20)
= ( ) [ ]

showing that as fracture length extends, the net treating pressure increases.
Eq. (20) for the PKN model and the similar relations in Table (4-1,4-2) for other models are very
illustrative of the "sensitivity" of fracture width to other parameters (viscosity and pump rate). For a
1
parameter sensitivity only to the 1/4 power, a 50% error results in only a 10% change (1.54 = 1.1) in the
calculated value so we see the calculations for fracture width are relatively insensitive to reasonable
uncertainties in these variables. This limited dependence on fluid viscosity is very important since
rheology of typical fracturing fluids (particularly gelled oil and cross-linked gelled water) are difficult to
characterize and measure, and exact rheology is sensitive to field mixing conditions. Thus there is
generally ALWAYS some uncertainty with respect to fracturing fluid viscosity.
Fracture Tip Effects, Fracture Toughness
While it was not emphasized, developing the relations between net pressure, , and height, pump rater
etc., in Table 4-1, implicitly assumed net pressure was "zero" right at the fracture tip. This is clearly not
the case since some non-zero pressure is required to break the rock and keep the fracture propagating. In a

7 PETE 4090-Unconventional Gas Reservoirs


simple fashion (and not quite mathematically correct as discussed below) the net pressure equation (for
the particular case of a confined height fracture, Eq. (20)) could be modified to include a "tip extension
pressure, , as seen below.
3
4 1
(21)
[ ]4 +

While not a rigorous mathematical relation, this equation gives a usable tool for comparing the relative
importance of fracture tip effects versus fluid-flow/ rheological effects. If fracture height is very large
(e.g., a radial frac) or if fracture length is very short (e.g.,GdK fracture geometry with L< H), then the
rheology effect on pressure is relatively small and tip effects become more important. Also, examining
equation [19], if Youngs modulus, E, is small, such as for very soft rocks like chalks and high porosity
(>30%) sandstone, then the rheology effect is small (since it is proportional to 1/E), and tip effects
become more important (or even dominant). For hydraulic fractures with some (not necessarily perfect)
height confinement in "normal" rock (e.g., modulus >1 to 2 106 psi), the rheology effect on net
pressure (and thus on fracture width) is much larger than the tip pressure, and tip effects become
insignificant. Empirically, pressure measurements on oil field scale hydraulic fractures usually show a tip
pressure or tip extension pressure of 100 to 200 psi.
As noted the equation above is not mathematically correct, e.g., the problem is more complex than simply
adding the net pressure (and frac width) due to viscous fluid flow to net pressure (and width) dale to a tip
extension pressure. The complications arise since a tip extension pressure increases width all along the
fracture, thus reducing the viscous flow pressure drop to some extent. This is discussed in detail by Nolte
[11] who also gives an analytical expression for wellbore pressure (and width) as a function of tip
extension and viscous pressure drop.
The magnitude of "tip extension pressure" (e.g., pressure at the fracture tip required to keep the fracture
propagating) is generally referred to in terms of the fracture toughness or strength, of the formation.
Actually, as briefly discussed below, actual processes occurring at the fracture tip are much more
complex than this, and the apparent fracture toughness for a hydraulic fracture is generally much larger
than a rock property" fracture toughness which might be measured in the laboratory.
Fracture toughness for materials is general defined in terms of one of two material properties: 1) critical
stress intensity factors (typical units of ) and 2) a surface energy released rate, G (with

typical units of ) [12]. For an elastic material, these two values are uniquely related to one another
2
by:
2
1 2 2
= = (22)

where E' is the plane strain models described earlier (Note that E' is essentially equal to Youngs modulus
with typical values from 1 107psi). Typical laboratory fracture toughness values were collected from
several published sources (for example, references [13] and [14]), with some of the data plotted in Fig.
(6).

8 PETE 4090-Unconventional Gas Reservoirs


For sedimentary rocks, may range from about 10 to moderately high values about
3000 with no clear dependence on rock type. Specificsandstone may have either higher or
lower fracture toughness than an adjacent shale layer[15]. The examples in the figure show a range of
values from about 1000 to about 3,000 , with a general trend of increasing (though not
dramatically) with Confining pressure. Though only four rock types are shown, this range is generally
good for a wide variety of rock types and conditions, with about2000 being a good, normal
number for the fracture toughness of a typical formation. However, as mentioned above, a propagating
hydraulic fracture often sees an apparent toughness which can be much larger than this.
As an example, consider a "penny shaped" or radial fracture. For such a fracture the critical net
pressure where the fracture starts to grow is found from [12]

= (23)
24
For a toughness of 2000 this gives a net propagation pressure of 72 psi (where any fluid
viscosity, fluid flow effects have been ignored) for a relatively small (by oil field standards) hydraulic
fracture with a radius of 50 feet (15 m), e.g., H = 100 feet.
However, as noted above fracture tip extension pressure is generally measured as 100 to 200 psi (or even
higher). This difference is the subject of much debate, but it is probably primarily related to an
unwetted region near to fracture tip as illustrated in Fig. (7). Very early in the history of hydraulic
fracturing, it was theorized [7] that the fracturing fluid could not reach the tip of propagating fracture.
Thus there will be a region ahead of the pressurized fracturing fluid where pressure inside the fracture is
(AT MAX) equal to reservoir pressure. For very low permeability formations, pressure in this region
could be a vacuum. The formation closure stress will then tend to close down the fracture in this region.
Retarding propagation and increasing the pressure required to force continued fracture growth (e.g.,
forcing a higher tip extension pressure). This phenomena has been well documented theoretically, and
has been clearly observed in both laboratory experiments [16] and semi- field scale mine back
experiments where hydraulic fracture growth was intensely monitored [17]. This phenomenon also
explains why bottomhole treating pressure for a main fracture treatment is generally exactly the same as
the pressure measured for a preceding mini-frac. The mini frac has already broken the rock, thus actual
fracture toughness is 0. In spite of the rock being already broken, there is seldom (if ever) any reduction
in subsequent injection pressure (ignoring the one time, often quite high, wellbore breakdown pressure).
While this fluid lag is a real phenomena, and attempts have been made to calculate or model the amount
of fluid lag (which has ranged from about 1% to 5% of fracture length in experiments), the exact lag will
be function of fluid viscosity, pump rate (e.g., fracture propagation rate), rock modulus, closure pressure,
and (very importantly) the pressure inside the unwetted" fracture tip region. This wide range of variables
makes calculating an apparent toughness for hydraulic fracture extremely difficult, and in practice, it is
necessary to measure this information in the field (e.g., direct measurement of fracture tip extension
pressure). Also since the actual rock fracture toughness plays little (it any) role in the behavior, lab tests
for fracture toughness are of little value in designing a hydraulic fracture treatment.
The role of fracture toughness (or apparent fracture toughness) is still under investigations, with recent
work being presented by Thiercelin [15], Jeffrey [18], and Gardner [19]. In general, all three concluded

9 PETE 4090-Unconventional Gas Reservoirs


that the role of toughness is more important for radial fractures, and that toughness played little (if any)
role for confined height fracture geometry. Their work might be summarized as seen in Table (4).
Rock Stress Effects
Just as fracture with results from the interaction of viscous flow and rock elasticity- the final fracture
geometry and height result from the interaction of fracturing pressure (e.g., net pressure. =
) with the properties and in situ stresses of the surrounding formations. While the mechanical
properties such as strength modulus, etc. may play a minor role, in general it is variations of in situ stress
(fracture closure stress) between formations that exercises the dominant control of fracture height growth
and fracture geometry.
This dominance is illustrated for the case of dolomite formations. Dolomite is among the hardest,
strongest, of rock types which are routinely hydraulically fractured; yet despite this strength and high
modulus, dolomite formations normally hydraulically fracture easily. This contradiction is related to the
chemical changes that transform limestone to dolomite. These changes cause a volume reduction in the
rock, and tor formations which are buried and surrounded by competent rock, this volume reduction
causes a decrease in the in situ confining stresses. Thus while the chemical changes create a strong, dense
rock, the stress reduction creates a situation where the formation is easily fractured.
The effects of rock stresses are quantified in chapters below on Rock Stresses and Fracture ; however, a
brief summary of in situ stress versus fracture geometry (or fracture height) is included here and is
illustrated in Fig. (8a-c).
When first initiated from a perforated (or open hole) interval a hydraulic fracture will assume an elliptical
shape as seen in Fig. (8a). The width of this fracture is governed by the "GdK model" (since fracture
height > length), and as fracture length increases, the treating pressure will decline as seen in Fig. (8c). As
pumping continues, the fracture will propagate into a circular or penny" shaped geometry since such a
radial geometry is the preferred natural shape for any hydraulic phenomena (e.g.; for a puddle of water on
a flat surface as well as for hydraulic fracture). This is illustrated as the fracture geometry at time-2 in Fig.
(8b). This radial fracture will continue to expand until a barrier to fracture growth is encountered, such as
at time-3 in Fig. (8) and during this growth the treating pressure continually declines.
Once a barrier to fracture extension is encountered, the fracture length begins to increase as seen in
Fig.(8b). For such growth, width is governed by the PKN model (since fracture length > height) and
treating pressure will begin to increase. Essentially during this stage of growth the fracture has become a
pipeline and as the pipeline becomes longer, pressure must increase in order to continue pumping down
the fracture at a constant rate.
Eventually, however, pressure becomes high enough to break out of the primary zone. Essentially, rock
formations have critical pressures, just as for any pressure vessel, and when this critical pressure is
exceeded, the formations break and fracture containment is lost. For the illustration, this occurs at 4 in
Figure (I-8b). At this point, fracture height growth will resume, however, fracture length will continue to
grow although at a reduced rate and net treating pressure will begin to level off at a value equal to the
critical pressure for the primary zone being fractured. If height growth causes the fracture to encounter
another low stress or easily fractured zone, then treating pressure will begin to decline as the fracture
preferentially grows into this new formation.

10 PETE 4090-Unconventional Gas Reservoirs


11 PETE 4090-Unconventional Gas Reservoirs
Reference
[1]. R. D. Carter, Derivation of the general equation for estimating the extent of the fractured area,
in: Appendix to Optimum fluid characteristics for fracture extension, by G. C. Howard and C.
R. Fast, Drill. and Prod. Prac. API. (1957).
[2].
[3]. Nolte, K.G.: "Determination of Fracture Parameters from Fracturing Pressure Decline," SPE 8341
presented at 54~1' An-al Fall Meeting, Las Vegas, Nevada, Sept. 23-26, 1 979.
[4]. Sneddon
[5]. Perkins, T.K. and Kern, L.R.: Widths of Hydraulic Fractures, J. Pet. Tech. (Sept. 1961) 937-
49; Trans., AIME, 222.
[6]. Geertsma, J. and de Klerk, F.: A Rapid Method of Predicting Fracture Width and Extent of
Hydraulically Induced Fractures, J. Pet. Tech. (Dec. 1969) 1571-81; Trans., AIME, 246.
[7]. Khristianovitch, S.A. and Zheltov, Y.P.: Formation of Vertical Fractures by Means of Highly
Viscous Liquid, Proc., Fourth World Pet. Cong., Rome (1955) Sec. II, 579-66.
[8]. Geertsma, J. and Haafkens, R., "A Comparison of the Theories to Predict Width and Extent of
Vertical Hydraulically Induced Fractures", presented at 31st Annual Petroleum Mechanical
Engineering Conference of the ASME, Mexico City, Sept 19-2 I, 1976.
[9]. Smith, M.B., Rosenberg, R. J., anal Gwen, J.F., "Fracture Width - Design Versus Measurement,"
SPE 10965, presented at the 57th SPY Fall Meeting, New Orleans, Louisiana, September 26-29,
1 9X2.

Appendix
PKN Model:

qi x f
w = 0.19[ ]1 / 4 (in field units, Newtonian fluid)
E'

w
wmax =
0.628

Non-Newtonian Fluid:
1 n '
2n'+1 n ' 0.9775 5.61 n ' 1 /( 2 n '+2 ) qi K ' x f h f 1 /( 2 n '+2 )
n'
128
wmax = 12[( )(n'+1)( ) ( )( ) ] ( )
3 n' 144 60 E

KGD Model: (in field units)

qi x f
2

wmax= w = 0.34[ ]1 / 4 ( )
E' h f 4

12 PETE 4090-Unconventional Gas Reservoirs

Potrebbero piacerti anche