You are on page 1of 4

Tech 101

Hydraulic Fracturing
To Improve Production

Ali Daneshy
Daneshy Consultants International

More about this topic can be found in Gidley, J.L., Holditch, S.A.,
Nierode, D.A., and Veatch, R.W. 1990. Recent Advances in Hydraulic
Fracturing. Monograph Series, SPE, Richardson, Texas 12.

Reasons for Hydraulic Fracturing


In their natural state, most oil and gas
wells do not produce at their optimum
level, but hydraulic fracturing can
address multiple challenges to efficient
production. Radial flow from the reservoir
into the wellbore is not an efficient
flow regime. As the fluid approaches
the wellbore, it has to pass through
successively smaller and smaller areas.
This causes jamming of the fluid and
reduction in flow. If one were to complete
the well such that the radial flow changes
to nearly linear, then the change in flow
pattern will increase well productivity.
A properly designed and executed
hydraulic fracture can change flow from
radial to nearly linear (Fig. 1). Nearwellbore permeability in most formations
is reduced by drilling, cementing, and
completion operations. Theoretically,
one can show that this permeability
reduction causes substantial reductions
in production rates (Fig. 2). Hydraulic
fracturing can extend the reach of the
wellbore beyond the damaged area and
abate its negative effect on production.
Hydraulic fracturing extends the reach
of the wellbore far into the formation.
Production is therefore controlled by
the properties of the average reservoir
reached by the fracture, instead of the
much smaller near-wellbore region. The
net effect is a reduction in the risk of
drilling into less- or nonproductive zones.
In formations with potential sand
production, hydraulic fracturing enables
financially viable production rates at
higher bottomhole flowing pressures.
This reduces the difference between
far-field and near-wellbore reservoir

14

pressures, and the tendency for sand


production. Production rates from most
wells eventually decline to a level below
which production is not economically
advisable. Hydraulic fracturing
increases the ultimate recovery factor
that corresponds to the economic cutoff
of production. For these and other
reasons, hydraulic fracturing is one of
the most common completion operations
in oil and gas reservoirs. Theoretically,
all wells can benefit from hydraulic
fracturing. However, the practice is
much more common in medium- and
low-permeability formations. In fact,
in many low-permeability reservoirs,
wells are fractured before production
is even attempted and is required to
exploit these resources economically.

Why and How Fracturing Helps


Production
The ability of a fracture to change the
flow regime and the distance the fracture
reaches into the reservoir determine
the production increase. Referring to
Fig. 1, if the reservoir fluid is to flow
into a fracture, the pressure inside the
fracture has to be close to the wellbore
pressure. This means that the hydraulic
fracture itself should offer very little
resistance to flow; in other words, it
should have very high permeability.
In reservoir engineering, this is
measured by the term nondimensional
conductivity, which is defined by:

CD=

k
k o X

where,
CD i s the nondimensional

fracture conductivity
ko is the formation permeability
k is the fracture permeability
is the propped fracture width
X is the propped fracture
length measured from the
wellbore to the fracture tip
A reasonable range for
nondimensional conductivity is
1< CD<10. For example, assuming
ko=1 md, =0.25 in., X =500 ft, and
CD =5, it is possible to compute a required
fracture permeability k =120 darcies.
Thus, a highly permeable fracture
needs to be created for a successful
treatment. This is not to say that a
less-permeable fracture will be
ineffective, but rather that a substantial
production increase requires a very
permeable fracture. As formation
permeability increases, the fracture
permeability required to achieve a
significant production improvement
becomes very large. On the other
hand, in high-permeability reservoirs,
the financial returns from even a
modest production increase can be
quite substantial. For example, a 10%
increase in a well producing 1,000 BOPD
means a 100 BOPD increase, while
a well producing 10 BOPD requires
a much larger percentage increase
in production to become financially
viable. Thus, financial and technical
successes in hydraulic fracturing
need to be reviewed separately on
the basis of different criteria.
At present, use of hydraulic fracturing
is much more prevalent in low- and
ultralow-permeability reservoirs. In
fact, most of these reservoirs would

Fig. 1Mechanics of production increase by hydraulic fracturing.

not even be under consideration


for development, were it not for
fracturing (e.g., the Bakken, Marcellus,
Haynesville, Eagle Ford, and many
other sandstones and shales). This is
particularly true in unconventional
reservoirs, where horizontal drilling
and the creation of multiple hydraulic
fractures along the length of these wells
have been enabling techniques for
the development of resource plays.

Mechanics of Hydraulic Fracturing


Fluid injection at high pressures
and rates causes initiation and
extension of hydraulic fractures.
The pressure required for fracture
initiation depends on the values of
the three in-situ principal stresses,
formation mechanical properties,
and the formations tensile strength.
The pressure required for fracture
extension is dominantly controlled by

1
0.9
0.8

Production Ratio

0.7
0.6
0.5

Damaged radius = 6.00 in.


Damaged radius = 9.00 in.

0.4

Damaged radius = 12.00 in.

0.3
0.2
0.1
0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Damaged-Zone Permeability/Formation Permeability Ratio

Fig. 2Effect of permeability damage on well productivity.

0.9

1.0

the least in-situ principal stress. Fracture


orientation is perpendicular to the
direction of the same principal stress.
During fluid injection, the fluid
pressure inside the fracture is higher
than the least in-situ principal stress, and
this keeps the fracture open. But after the
injection stops and pressure is allowed
to drop, the fracture begins to close. To
keep the fracture open and conductive
(permeable) after the treatment, a
proppant is mixed with the fluid and
injected inside the fracture to keep it
open during production operations.
Fig. 3 shows a fracturing chart with
various stages of the treatment marked
on it. The vast majority of fracturing
treatments use water as the base fluid
and add to it various chemicals to
give it specific physical and chemical
properties. Each fracturing job
begins with injection of a prepad,
which usually consists of a mixture of
mid- to low-strength acid and water.
This is followed by a pad, which is a
mixture of water and a viscosifier or
friction reducer (usually a polymer).
This is then followed by the slurry
which is a mixture of proppant and the
fracturing fluid. The specific details
of fluid mixture and proppant type
vary between reservoirs. Once the
desired amount of proppant is pumped,
in the last step, the slurry inside the
wellbore is displaced into the fracture.

Fracturing Materials
Fracturing materials consist of two
main components: fluid and proppant.
The most common fracturing fluid
is a mixture of water and additives.
Common additives include polymerbased viscosifiers (to increase fluid
viscosity to carry the proppant inside
the fracture), friction reducers (to reduce
friction pressure of fluid moving inside
the wellbore), breakers (to break
the fluid and reduce its viscosity after
the treatment, so it can flow back),
clay stabilizers (to eliminate damage
to swelling clays that may exist in the
formation), biocides (to prevent fluid
degradation by bacteria), buffers
(to adjust fluid pH), surfactants, and
nonemulsifiers. In some instances, water
is mixed with nitrogen or carbon dioxide
to form a foam. Oil-based fluids, such

Vol. 6 // No. 3 // 2010

15

5000

Pad

40
35
30

Slurry

Pressure, psi

4000

25

3000

20
Pressure

2000

15

Slurry Rate
Proppant Concentration

10

1000

0
0

20

40

60

80
100
Elapsed Time, min

120

140

0
160

Slurry Rate, bpm; Proppant Concentration, lb/gal

Displacement

Prepad

6000

Fig. 3A typical fracturing chart illustrates the steps to hydraulically fracture a well.

as kerosene, diesel fuel, or propane are


also occasionally used in formations
that are highly water sensitive.
The most common proppant is natural
sand that meets very specific physical,
mechanical, and chemical requirements.
These include size (commonly varying
from 100 mesh to as large as 10/20
mesh), sphericity, mechanical-crush
strength, and very low solubility in acid.

Synthetic proppants are used for special


applications, such as deep formations.
The more common compositions
are ceramics, bauxite, walnut hulls,
or even plastics (the latter two are
lightweight and used mostly with very
low-viscosity fluids). Fig. 4 shows some
of the common industrial proppants.
As stated previously, one of the
methods of production enhancement is

Ottawa

Low Density

Resin Coated

Bauxite

Fig. 4Common industrial proppants used in


hydraulic fracturing to maintain fractures.

16

to create a high-permeability fracture


that extends deep into the formation.
To achieve this, the proppant needs to
be carried far along the fracture. Since
the specific gravities of most industrial
proppants are between 1.9 and 3.6,
(much higher than water), the main
mechanism for carrying proppant inside
a fracture is fluid viscosity. Common
fracturing fluid viscosities range from a
few tens to several thousand centipoises.

Fracture Design
Engineering computations always
precede a fracturing treatment. These
consist of calculation of fluid volume
and viscosity, injection rate, weight of
proppant, volumes of different phases
of the job (prepad, pad, slurry, and
displacement), surface and bottomhole
injection pressure, hydraulic horsepower
required at the surface, and the
mechanical equipment needed for this.
A very important part of fracture
design is determination of the fluid
volume required to create a fracture
with a given length. A hydraulic
fracture is usually identified by three
dimensions: length, width, and height
(Fig. 5). Fracture length itself has two
components: created and propped.
Created length is the distance between
the wellbore and farthest point into the
formation. Propped length is the distance
between the wellbore and farthest point
where proppant has travelled inside the
fracture. Fracture width is the separation
between the two faces of the fracture.
Its value is largest at the wellbore and
tapers toward the tip of the fracture.
Fracture height is the distance between
the top and bottom of the fracture.
The relationships between these
parameters and fracture design are
as follows. (1) Created fracture length
influences total injected volume and
fracture width. The longer the required
fracture length, the larger the volume of
fluid needed to create this length. Longer
length also results in a wider fracture. (2)
Propped fracture length influences slurry
volume, proppant weight, and production
increase. Creating a longer propped
fracture length requires injecting a
larger amount/weight of sand. But this

The Future: Fracturing


Horizontal Wells
In recent years, hydraulic fracturing
technology has witnessed major
advances, as its use has been the main
driver for production from tight oil and
gas reservoirs in the US and Canada. In
these applications, numerous hydraulic
fractures are created along the length
of an open or cased horizontal well to
gain access to large volumes of the lowpermeability reservoir. The number of
fractures in cased horizontal wells can
be as many as 60 to 70. This number is

smaller (but still steadily increasing)


in open holes. Most treatments in gas
wells are done with water plus a friction
reducer pumped at very high rates.
In tight oil reservoirs, the fluid usually
has higher viscosity and is injected at
lower rates. In both applications, the
volume of fluid and weight of proppant
is at least an order of magnitude larger
than what is used in vertical wells.
Hydraulic fracturing has played a very
important role in enabling the oil and
gas industry to meet the energy needs
of the growing world. As we are forced
to rely on more marginal reservoirs for
meeting our energy requirements, the

Height

also results in a higher production


increase. (3) Fracture width depends
on formation mechanical properties,
fluid viscosity, and fracture dimensions.
Higher-viscosity fluids create a wider
fracture. The longer the fracture, the
wider it will be at the wellbore. (4)
Fracture height is the big unknown
and is usually assumed to be constant
and related to formation thickness.
The relationship between fracture
width and other fracture dimensions
has been a subject of debate among
fracturing experts for many years. The
two basic concepts are those of Perkins
and Kern (1961), who relate fracture
width to its height, and Khristianovic
and Zheltov (1955), who express
fracture width as a function of its length.
Depending on the choice of basic
equations, elaborate computations are
needed to calculate various fracturing
parameters as functions of injected
fluid and formation properties. These
are usually done using computer
simulations, which are commercially
available for this purpose.

need for better fracturing technology


becomes even more pressing. This is
best demonstrated by the recent rapid
development and deployment of novel
fracturing tools and techniques that are
specifically suited for horizontal wells.
References
Khristianovic, S.A. and Zheltov, Y.P. 1955.
Formation of Vertical Fractures by Means of
Highly Viscous Liquid. Proc., Fourth World
Petroleum Congress, Rome, Sec. II, 579586.
Perkins, T.K. and Kern, L.R. 1961. Widths
of Hydraulic Fractures. J. Pet. Tech. 13
(9): 937949. doi: 10.2118/89-PA

Length

Fig. 5Dimensions of a hydraulic fracture relative to the wellbore.

Ali Daneshy is president of Daneshy Consultants International, where he provides fracturing


consulting and training services, and adjunct professor at the University of Houston, where he
teaches a graduate course on hydraulic fracturing. He has more than 40 years of global experience
in the technology and operations of hydraulic fracturing. He is the recipient of the SPE Distinguished
Member and Distinguished Service Awards for his contributions to hydraulic fracturing, including
more than 40 technical publications. Daneshy is a very active member of SPE and a past member
of its Board of Directors. He received an MS degree from the University of Minnesota and a PhD
degree from the University of Missouri-Rolla, both in mining engineering (rock mechanics).

Vol. 6 // No. 3 // 2010

17

You might also like