You are on page 1of 22

SPE-174858-MS

A Three-Dimensional Thermal Model for Hydraulic Fracturing


Kaveh Amini, Mohamed Y. Soliman, and Waylon V. House, Texas Tech University

Copyright 2015, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Annual Technical Conference and Exhibition held in Houston, Texas, USA, 28 –30 September 2015.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents
of the paper have not been reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect
any position of the Society of Petroleum Engineers, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written
consent of the Society of Petroleum Engineers is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may
not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
Distributed Temperature Sensing (DTS) can provide the wellbore temperature profile during the hydraulic
fracturing treatment. This temperature profile is a complex function of many parameters including leakoff
coefficient, thermal conductivities of the rock and the injected fluid, minimum stress, porosity, etc.
Because of these complexities analytical solutions are not capable of adequately determining the effect of
these various parameters on the temperature profile. Moreover, these parameters may vary from one layer
to another. Therefore, a three-dimensional temperature model is needed to estimate the fracture param-
eters by matching the wellbore temperature profile with simulation results.
A three-dimensional temperature model was developed to determine the temperature profile in and
around the fracture based on a pseudo three-dimensional (P3D) fracture propagation model. The propa-
gation model, the wellbore thermal model, and the fracture thermal model were solved simultaneously to
consider the effect of temperature and pressure on various parameters, such as specific volume. Moreover,
the results of the fracture propagation simulation of five cases were compared with the results of full 3D
simulators, obtained from literature.
The fracture half-length, fluid efficiency and the net pressure at the entrance of the fracture simulated
by this model were in good agreement with the results of full 3D simulation. The minifrac propagation,
closure and its temperature were also simulated by this model. The pressure decline curve was predicted
by simulator while the effect of temperature and pressure on other variables, such as fracturing fluid
density and viscosity, was considered. Additionally, it was observed that the location of optical fiber cable
was very important in DTS temperature measurement. Accordingly, a method is provided to consider the
effect of the distance of the fiber optics cable from the fracture.
The development of this hydraulic fracturing 3D thermal model is another step for the interpretation
of the wellbore thermal profile, measured by the DTS technology during hydraulic fracturing treatment.
This model is also capable of being used for more accurate hydraulic fracturing design and analysis.

Introduction
The temperature data are valuable information for monitoring a reservoir. Furtunately, distributed
temperature sensing (DTS) technology makes it easier to obtain the temperature data along the wellbore.
However, DTS cannot measure the temperature of events happening far from the wellbore; for example,
the temperature of the hydraulic fracturing fluid traveling along the fracture cannot be measured. In fact,
2 SPE-174858-MS

we need to combine the fracture propagation model and heat transfer model to estimate the temperature
inside and around the fracture. The estimated temperature at wellbore could be compared with the
measured data to match the uncertain fracture parameters.
Since many parameters affecting fracture propagation are temperature dependant, the combination of
heat transfer model and fracture propagation model is essential for accuracy of both models. However,
only few works have been published regarding this combination. Among these works, Meyer (1989)
presented an analytical method, coupling thermal and 2D fracture propagation models. Seth, Reynolds,
and Mahadevan (2010) also combined 1D fracture propagation and thermal numerical models to interprete
DTS data.
Our model consists of a P3D fracture propagation model, the fracture and the rock 3D thermal models,
and the wellbore thermal model, which are solved simultaneously. This model can be used to interprete
DTS data during hydraulic fracturing, to simulate hydraulic fracture temperature, and to design and
analyse hydraulic fractures in complex conditions.

Pseoudo-3D Propagation Model


Continuity Equation
To make it simple, let us consider a vertical fracture propagating through a horizontal payzone. The
payzone is contained between upper and lower barriers. The final model can be applied for multiple
payzone and several barriers. The diffential volume element is a vertical slice of the fracture which is
prependicular to the fracture propagation direction. The continuity equation around the differential
ellement can be expressed as
(1)

where A is the cross-sectional area of the fracture, Q is the volume flow rate passing through the
fracture cross-section, and ql is the leakoff volume flow rate per unit of the fracture length in each side
of the fracture. The numerical format of the continuity equation around the element is written as
(2)

Figure 1—A fracture element for P3D propagation model

The fracture cross-section area can be calculated from


SPE-174858-MS 3

(3)

To numerically calculate cross-sectional area, each vertical element is divided to sub-element with
variable width, , and constant height, ⌬z. Therefore,
(4)

where k is the index of sub-elements which changes along z-axis, Nhl,i and Nhu,i are the minimum and
the maximum values of k-index, respectively.
The fluid-loss or leakoff coefficient is an important parameter in hydraulic fracturing treatment design
which significantly affects the success of the fracturing treatment. Carter (1957) developed an equation
for leakoff velocity:
(5)

where ␯l is the apparent leakoff velocity, Ct is the leakoff coefficient, t is the time elapsed since the
initiation of the fracture, and ␶(x,z) is the arrival time of the fracture tip at the coordinate (x,z). By
integrating from Eq. 5, one can obtain the average leakoff velocity during the time step ⌬t:
(6)

The leakoff volume flow rate per unit of the fracture length is obtain from
(7)

which can be converted to numerical format as


(8)

Spurt loss Vsp is considered as a one-time sudden leakoff volume per unit of fracture surface, affects
on the surface area Asp at the time t⫽␶. Vsp is assumed to be a constant value for a specific rock and fluid
in this model. V is the sudden leakoff volume on the fracture element’s surface between the time steps
nt-1 and nt, and can be calculated by
(9)

Since the spurt-loss effect is observed during the time period ⌬t, it can be expressed as an extra leakoff
volume flow rate per unit of the fracture length, qsp:
(10)

Pressure gradient inside the fracture


To calculate the pressure gradient inside the fracture, the fracture cross-section is considered as a
combination of several slit-flow ducts. The potential function p’ is related to the volume flow rate through
the following equation (Nolte 1979):
(11)

where
(12)

(13)
4 SPE-174858-MS

(14)

Eq. 11 is valid for isothermal steady uniform laminar flow of power-law fluid (Kozicki, Chou, and Tiu
1966). K is the fluid consistency index, and n is the flow behavior index. It is assumed that the potential
function is constant in each vertical cross-section. Eqs. 11 through 14 can also be presented as
(15)

(16)

(17)

Let us consider the same fracture elemet as illustrated in Fig. 1. Eq. 11 can be applied for the element
as
(18)

where is the fracture width at the fracture elemet center. and are obtained from
(19)

(20)

Fracture Geometry
The stress intensity factor KI is a measure of the fracture energy for propagation. It is a function of the
fracture geometry and the net pressure. If the stress intensity factor becomes larger than its critical value
KIC, which is also known as fracture toughness, the fracture propagates. The stress intensity factor can be
calculated by the following equations when the fracture propagates through some layers that have
different effective horizontal stress and fracture toughness (Valkó and Economides 1995):
(21)

(22)

where KI⫹ and KI- are the stress intensity factor for the upper and the lower fracture tip, pn is the net
pressure, and h is the fracture height. Eqs. 21 and 22 are utilized to numerically calculate the fracture
height.
Under assumption of linear elasticity, the fracture opening is calculated using the method presented by
Morales and Abou-Sayed (1989).

Considering Compressibility and Thermal Expansion


The temperature and the pressure of the fracturing fluid change from the wellhead to the downhole. They
are even different at the fracture entrance and at the fracture tip. It is reported that ignoring compressibility
and thermal expansion results in major errors in predicting leakoff coefficient using pressure decline
SPE-174858-MS 5

method (Soliman 1986). Therefore, the compressibility and the thermal expansion are considered in the
fracture propagation and closure model. The compressibility is defined by
(23)

and the thermal expansion is defined as


(24)

where ␯ is the specific volume, p is the pressure, and T is the temperature. The specific volume can
be calculated by integrating of Eqs. 23 and 24:
(25)

where po is the reference pressure, To is the reference temperature, and ␷o is the specific volume at the
reference pressure and temperature. Eq. 25 is valid if the compressibility and the thermal expansion are
constant between po to p and between To to T. Alternatively, the specific volume can be experimentally
obtained in the lab and imported into this simulator as a table.
The flow rate Q is a function of the flow rate in the reference conditions, Qo:
(26)

which can be substituted in the continuity equations (Eqs. 1 and 2) and the friction pressure drop
equations (e.g. Eq. 15). The density is also modified by following equation:
(27)

where ␳o is the fracturing fluid density at the reference pressure and temperature.

Precalculating Fracture Parameters as a Function of Pressure


In this Pseudo 3D propagation model, it is assumed that the pressure is constant in each vertical cross
section. Considering this assumption, many parameters such as fracture width, fracture height, and
fracture cross-sectional area are functions of the pressure (or net pressure). Even the parameter R in Eq.
15 is a function of pressure.
In order to decrease the simulation time, a table of different parameters versus pressure is precalcu-
lated. So during the simulation, the simulator only reads the parameters from the table and performs an
appropriate interpolation.

3D THERMAL MODEL
The thermal model consists of three parts: 1- thermal model around the fracture, 2- thermal model in the
rock matrix, and 3-welbore thermal model.

Temperature Model around the Fracture


Let us consider a narrow vertical slice of a fracture as presented in Fig. 2, and write energy balance
equation around this element as bellow:
(28)
6 SPE-174858-MS

Figure 2—Vertical slice of fracture as an element in heat transfer model and energy transfer around the element

As illustrated in Fig. 2, the fracturing fluid is injected into the element at flow rate Qi and temperature
Ti-1. The fluid either flows out of the element along the fracture at rate Qi⫹1 and temperature Ti or leaks
off into the rock matrix at rate QL and temperature Ti.
Fig. 2 illustrates the details of energy transfer around the element. The items 1 to 8 in Fig. 2 are
described as below:
1- Input Energy and Work due to the in-coming flow along the fracture
(29)

(30)

(31)
where vi and Ti are the inward flow velocity and the temperature of the i-th element.
2- Output Energy and Work due to the out-going flow along the fracture
(32)

(33)

(34)

3- Output Energy and Work due to the Leakoff


(35)

(36)
The leakoff flow rate around the fracture element is not uniform; for example, the area close to the tip
may have recently propagated and have the higher rate of leakoff. Therefore, each element is divided to
sub-element with the height ⌬z.
SPE-174858-MS 7

(37)

where vy is the apparent leakoff velocity and ⌬S is the sub-element’s side surface area:
(38)

4- Energy Storage
(39)

(40)

5-Work Performed to Move the Rocks


(41)

6- Heat Conduction from Walls into the Fluid Since the fracture is narrow and the velocity of the
fluid in the fracture is low, we can assume that the heat transfer mechanism between the fracture walls
and the fracturing fluid is mainly heat conduction:
(42)

where Ti is the temperature of the i-th fracture element, and Tw is the temperature of an adjacent rock
block. The heat transfer between the fracture and the walls is symmetric, so as illustrated in Fig. 3, we
assume that there is an imaginary wall in the middle of fracture which prevents heat and fluid flow. It
divides the fracture block into two half-blocks. ⌬yw is defined as the distance from center of the fracture
half-block (point A in Fig. 3) to the center of adjacent block on the wall (point B):
(43)

Figure 3—Schematic diagram of the heat transfer between the fracturing fluid and the walls

where w៮ i is the average width of i-th element, and ⌬y is the wall blocks width. w៮ i is obtained from
(44)

Ai, Ai⫹1, Hi, and Hi⫹1 are illustrated in Fig. 2. k is overall heat conductivity between points A and B
in Fig. 3. k is calculated by
8 SPE-174858-MS

(45)

Applying Eqs. 43, 44, and 45 into Eq. 42 results in


(46)

7- Heat Conduction through Fluid into the Fracture Element


(47)

8- Heat Conduction through Fluid out of the Fracture Element


(48)

Energy balance inside the fracture If the input, output, and stored energy and work are replace in Eq.
28, the energy balance around a vertical element of the fracture will be obtained as
(49)

Cw and Ck are unit conversion coefficients:

By substituting Eqs. 29 through 48 into Eq. 49, the following equation will appear:
(50)

where
(51)

The superscript n and n⫹1 represent the time steps. If nothing is mentioned, the time step is n⫹1. The
function ⍀ represents the overall effect of kinetic energy and work.
3D Thermal Model in Rock Matrix
It is assumed that the leakoff fluid moves only in y-axis direction (perpendicular to fracture surface). It
is also assumed that the leakoff fluid sweeps the formation piston-wisely. To simplify the model, it is
assumed that the rock matrix and the fluid inside the rock pores have the same temperature at any location.
Therefore, the differential heat balance equation inside the rock matrix is simplified to
(52)

where k is the effective heat conductivity and M is the effective volumetric heat capacity:
SPE-174858-MS 9

(53)

(54)

Numerical Solution of the Rock Matrix Thermal Model At the block (i,j,k), inside the rock matrix,
(55)

(56)

(57)

where . We can write equations similar to Eq. 57 for and . The

rock matrix thermal model for the uniform ⌬y, ⌬z, and ⌬t can be simplified to
(58)

During the solution of the thermal model, the calculation time increases dramatically with the decrease
of the gridblock size. Therefore, the gridblock size in thermal model is recommended to be choosen larger
than in propagation model.
The Boundary Conditions for the Rock Matrix Thermal Model Depending on the position of the
boundary, it is approximated with either 1- no flow boundary due to symmetry, 2- very far boundary, or
3- locating on fracture walls.
For example, at j⫽Ny, the very-far boundary condition governs:
(59)

Therefore, the heat balance equation (Eq. 52) reduces to


(60)

There are six boundary conditions on the faces of grid system, and twelve boundary conditions on the
edges which are the combination of face boundary conditions. There are also eight boundary conditions
on vertexes.

RESULTS AND DISCUSSION


Case Studies
In order to test the validity of our P3D propagation model, five case studies were selected from Ref.
(Weng 1992). These five cases had been simulated by the same reference using the full-3D simulators
TerraFrac™ and University of Texas (in this paper, they are called as 3D simulator A, and 3D simulator
10 SPE-174858-MS

B), and a modified P3D model. All cases were simulated by our fracture propagation simulator, and
presented in Table 4 to be compared with the results of full 3D and modified P3D models.
The input data, as they mentioned in Ref. (Weng 1992), needs some processes, such as unit conversion,
to be applicable in our simulator. Some more assumption is taken; for example, critical stress intensity
factor of the pay zone is taken as zero. Moreover, some variables related to the flow pattern inside a
fracture, such as “the number of perforations”, or “the perforated interval”, are not considered. The
modified data was written in Table 2.
Cases 1 and 2 are symmetric fractures using Newtonian fracturing fluid. Both cases are almost the same
except that in Case 2, there is no leakoff in barriers. There is no spurt loss in either of cases. Cases 3 and
4 are also symmetric. The fracturing fluids are non-Newtonian in both cases; however, they have different
consistency and behavior index. The most important different between these two cases is that there is a
lower stress contrast, 100 psi, in Case 3 comparing with 500 psi stress contrast in Case 4. The leakoff
coefficient and the spurt loss are uniform within all layers including the pay zone and the barriers. The
simulated fractures of Cases 3 and 4 are shown in Fig. 4 and Fig. 5, respectivvely. The dark blue color
in these figures means the zero fracture width, and the dark red means the maximum width. Case 5 is a
non-symmetric fracture with the stress strain of 900 psi and 1400 in upper and lower barriers, respectively.
The fracturing fluid is non-Newtonian (power-law). The leakoff coefficient and the spurt loss are lower
than in Cases 3 and 4, and they are uniform within all layers including the pay zone and the barriers. The
critical stress intensity factor of the barriers in Case 5 is much higher than in other cases. The critical stress
intensity factor of the pay zone is considered zero in all cases, which means that the pay zone is
completely brittle.

Figure 4 —Simulated fracture of case 3 with low stress contrast

Figure 5—Simulated fracture of case 4 with high stress contrast

The comparison between the results of different models shows that our fracture propagation model is
in a good agreement with other models on the net pressure at the wellbore, the fracture half-length and
the fracturing slurry efficiency. However, it seems that the fracture height is overestimated by our model
SPE-174858-MS 11

due to the simplistic assumption of “1D fluid flow inside the fracture”. This assumption and the calculated
fracture height at wellbore is acceptable in open-hole fracturing treatment, while in cased-hole fracturing
treatment there is a 2D or even 3D fluid flow pattern within the fracture.

Net Pressure at Fracture Entrance


Fig. 6 shows how the net pressure at the fracture entrance increases during propagation. The horizontal
axis is the time in logarithmic scale and the vertical axis is the net pressure. It is clear that the logarithmic
regression of net pressure against elapsed time is a very good match for simulated data. Therefore, during
fracture propagation the following equation is valid:
(61)

Figure 6 —Net Pressure at the entrance of fracture during fracture propagation for cases 1 to 5

This equation is very helpful to make a good guess for net pressure at the fracture entrance during
propagation, and dramatically decreases the number of iterations. If p1 and p2 are the net pressures at times
t1 and t2, the net pressure at the fracture entrance for the next time step (tn) is approximated by
(62)

Combined Propagation and Temperature simulation


Let us define Case 6 with the input data mentioned in Table 2 and Table 3. A grids system of
500x600-block is used in the propagation and the closure simulation. Each block is 1 ft length and 0.4 ft
height. The thermal model has one more dimension which is perpendicular to the fracture surface (y-axis),
and use larger blocks, which are 50 ft length, 20 ft height, and 0.2 ft width
Fig. 7 shows the simulated shape of the fracture at the end of pumping period. The dark blue color
represents the zero fracture width, and the dark red color represents the maximum width, which is 0.47
in. A summary of the fracture propagation simulation is presented in Table 5.
12 SPE-174858-MS

Figure 7—Case 6 Simulated fracture at the end of pumping. The color map on the left shows the fracture width.

.
Fig. 8 illustrates the fracture wall temperature at the end of pumping. Although the fracture half-length
is 407 ft at the end of pumping, the fracture wall temperature remains unchanged 200 ft away from the
wellbore. The fracturing fluid warms when it is flowing inside the fracture and becomes almost as warm
as the formation at 200 ft away from the wellbore.

Figure 8 —Temperature of the fracture walls of case 6, at the end of pumping

Leaked-off Fluid Front In order to know how far the leaked-off fluid penetrates, the leaked-off fluid
fronts in a horizontal plane, located at the center of pay zone, were plotted in Fig. 9 at different times. The
horizontal line is the distance swept as the result of spurt loss. The pumping ends after 50 minutes, but
the leakoff continues 52 more minutes when the fracture takes the final shape and stops closing. So in
Case 6, the leakoff fluid front will not move forward after 102 minutes that can be seen in Fig. 9 where
the 100-minute and 125-minute fronts are almost located on each other.
SPE-174858-MS 13

Figure 9 —The front of leakoff fluid at the center of payzone, at different times (case 6)

It can be observed that in this case, the spurt loss has a considerable share in the total leakoff. It sweeps
0.4 inch of the formation. The 102-minute leakoff, which includes pumping and closure, sweeps only a
few inches (about 1.5 inches) of the formation. The farthest point that the leaked-off fluid penetrates is
1.7 inches.
The distance swept by leakoff fluid is plotted versus square root of time in Fig. 10. The simulated data
should locate on a straight line with the slope of 2Ct/␾. The result completely matches with this amount,
which reaproves the accuracy of this model.

Figure 10 —Distance swept by leakoff fluid: A method to verify the validity of the simulation

Formation Temperature Fig. 11 shows the formation temperature in a horizontal plane, located at the
center of pay zone, at the end of pumping of Case 6. The temperature is shown by different colors which
are described by color-bar at the right side of the figure. These diagrams show that the temperature of the
formation farther than 0.6 feet from the fracture remains unchanged during 50 minutes of fracturing
treatment. In other words, the temperature gradient is very high in y-direction.
14 SPE-174858-MS

Figure 11—Formation temperature at the center of payzone, at the end of pumping of case 6, illustrated in 2D diagram

Location of DTS Optical Cable It was mentioned that the temperature gradient is high in y-direction, so
the location of DTS cable affects the measured temperatures. We are trying to observe the possible DTS
measurements if it locates in different distances from the fracture surface, and finally trying to estimate
the distance between DTS cable and the fracture. The presented data is related to Case 6.
Four different locations for DTS cable were considered: 0.1 ft, 0.3 ft, 0.5 ft, and 0.7 ft away from
fracture. Fig. 12 shows how we would measure the temperature in each scenario. It can be seen that the
temperature profiles are completely different from each others. Therefore, the distance between the
fracture and the cable should be considered as a parameter to find a good thermal match.

Figure 12—Formation temperature profile in different distances to the fracture in case 6

Fortunately, there is a clue to figure out the distance between the cable and the fracture. When the
pumping stops, the formation temperature starts rising. However, this temperature-rise occurs after a
time-lag which depends on the distance between the cable and the fracture. In Fig. 12, the minimum
SPE-174858-MS 15

temperatures which are indicated by a triangular marks and dash line are observed in different time-lags.
Table 1 shows the minimum temperature and the time-lage to observe minimum temperature for different
location of DTS cable. The time-lag also depends on the formation effective thermal conductivity,
effective volumetric thermal capacity, and the leakoff rate.

Table 1—The minimum temperature and the time-lag to observe minimum temperature for different location of DTS cable

3D Tempermal Model of a Minifrac


Let us consider a minifrac case (Case 7) similar to case 6 except that the minifrac pumping duration is 15
min, the cool-down duration is 135 min, the grid size length is 250 ft, and the block size of thermal model
is 25 ft.
Fig. 13 shows the simulated wellbore pressure of the minifrac. After the pumping stops, the pressure
becomes uniform inside the fracture. Therefore, the fracture narrows, and may continue propagating while
the fracturing fluid leaks off to the formation, and the pressure declines at the wellbore. In Case 7, the
fracture length pumping is 181.5 ft at the end of pumping as shown in Fig. 14. The minifrac continues
propagating to 215 ft 13 minutes after shut-in, as demonstrated in Fig. 15. The minifrac continues
propagating up to 217 ft 23 minutes after shut-in.

Figure 13—Simulated wellbore pressure of the minifrac


16 SPE-174858-MS

Figure 14 —The fracture surface temperature of minifrac case 7, at the end of pumping

Figure 15—The fracture surface temperature of minifrac case 7, after 15 min pumping and 13 min shut-in

Conclusions
● The Pseudo Three Dimensional (P3D) fracture propagation model was improved. Although P3D
model had been developed decades ago, the model details and the numerical method were
improved to simulate more accurate results in a shorter time.
● In P3D model, the fracture height and the fracture width at each point are functions of net pressure
at that point. Therefore, this simulator calculates a table of fracture height and width versus net
pressure. So, during several iterations, the simulator reads the amount of fracture height and width,
instead of their calculation. This technique incredibly reduces the simulation time.
● A three dimensional thermal models was developed to determine the temperature in the fracture
and in the formation.
● The fracture propagation, closure, and thermal models were numerically solved togheter to provide
a fracture thermal simulator. This simulator considers the effect of pressure-dependant or tem-
perature-dependant properties.
● The simulator accepts various rock properties, leakoff coefficient, and spurt loss for different rock
layers. It also accepts various fracturing fluid rate and properties as the function time.
● The results of simulation of five cases were compared with the results of full 3D simulators,
obtained from literatures. The fractures half-length, fluid efficiency and the net pressure at the
entrance of the fracture were in a good agreement with the results of full 3D simulation. However,
the fracture height was overestimated due to the assumption of 1D fracturing fluid flow.
SPE-174858-MS 17

● The results of simulations show that the net pressure at the entrance of a fracture is a logarithmic
function of time during the pumping period. This concept is applied to make a good initial guess
for the net pressure at the entrance of a fracture at each time step.
● It was observed that the location of optical fiber cable was very important in DTS temperature
measurement. The distance between the cable and the fracture can be figured out by the time-lag
between the end of pumping and the time when the minimum temperature was measured.

Nomenclature
A Cross-sectional area of fracture, ft2
Asp Area of fracture face affected by spurt loss, ft2
c Heat capacity, BTU/ft3.F
cl Heat capacity of pore fluid, BTU/lbm.F
cr Heat capacity of rock matrix, BTU/lbm.F
Ct Leakoff coefficient, ft/sec½
Cw, Ck Unit conversion coefficients
Econd Heat conducted through fluid along the fracture, BTU
Eh Heat conveyed by flow, BTU
Ek Kinetic energy of flowing fluid, lb.ft2/sec3
Ew Heat conducted from wall into the flowing fluid, BTU
G Shear Modulus
h Fracture height, ft
K Fluid consistency index, (lbf.secn)/ft2
k Overall thermal conductivity between fracture fluid and rock, BTU/ft.sec.F
keff Effective thermal conductivity, BTU/ft.sec.F
KI- Stress intensity factor for the lower fracture tip
KI⫹ Stress intensity factor for the upper fracture tip
kl Thermal conductivity of pore fluid, BTU/ft.sec.F
kr Thermal conductivity of rock matrix, BTU/ft.sec.F
kx, ky, kz Effective thermal conductivity in x, y, or z direction, BTU/ft.sec.F
M Effective volumetric heat capacity, BTU/ft3.F
n Flow behavior index
Nhl,i Minimum values of k-index for i-th element
Nhu,i Maximum values of k-index for i-th element
p Pressure, psi
p’ Potential function, psi
pn Net pressure, psi
po Reference pressure, psi
Q Volume flow rate passing through fracture cross-section, ft3/sec
QL Leakoff volume flow rate, ft3/sec
ql Leakoff volume flow rate per unit of fracture length, ft2/sec
Qo Flow rate at reference conditions, ft3/sec
⌬S Sub-element side surface area, ft3
t Time, sec
T Temperature, °F
To Reference temperature, °F
Tw Temperature of fracture wall, °F
V Volume, ft3
18 SPE-174858-MS

V Volume of fracture element, ft3


v៮ l Average leakoff velocity, ft/sec
vl, vy Apparent leakoff velocity, ft/sec
Vsp Spurt loss
w Fracture width, ft
W Work, psi.ft3/sec
Wi Work done on i-th element by inward flow pressure, psi.ft3/sec
w៮ i Average width of i-th fracture element, ft
wi,k Width of sub-element (i,k), ft
WL Work gets out of i-th element by leakoff flow pressure, psi.ft3/sec
x Lateral coordinate along fracture length, ft
y Coordinate perpendicular to fracture face, ft
⌬y Wall blocks width, ft
z Vertical coordinate, ft
⌬z Height of sub-element, ft
␣ Thermal expansion, °F-1
␤ Compressibility, psi-1
␯ Specific volume, ft3/lbm
␯o Specific volume at the reference pressure and temperature, ft3/lbm
␳ Fracturing fluid density, lbm/ ft3
␳l Density of pore fluid, lbm/ ft3
␳o Fracturing fluid density at reference pressure and temperature, lbm/ ft3
␳r Density of rock matrix, lbm/ ft3
␶ Arrival time of fracture tip at the coordinate (x,z), sec
␾ Porosity
⍀ Overall effect of kinetic energy and work, Defined by Eq. 51

Subscripts
ave Average value in fracture element
c Vertically center of fracture element
i Index of fracture elements which changes along x-axis
j Index of rock grid blocks which changes along y-axis
k Index of fracture sub-elements which changes along z-axis
L Leakoff flow

Superscripts
nt Time step index in thermal model
t Time step index in fracture propagation model

References
Carter, RD. 1957. Derivation of the general equation for estimating the extent of the fractured area (in
Appendix I of “Optimum Fluid Characteristics for Fracture Extension,” Drilling and Production
Practice, GC Howard and CR Fast, New York, New York, USA, American Petroleum Institute:
261–269.
Kozicki, W., C. H. Chou, C. Tiu. 1966. Non-Newtonian flow in ducts of arbitrary cross-sectional
shape (in Chemical Engineering Science 21 (8): 665–679.
SPE-174858-MS 19

Meyer, Bruce R. 1989. Heat transfer in hydraulic fracturing (in SPE Production Engineering 4 (04):
423–429.
Morales, RH, AS Abou-Sayed. 1989. Microcomputer analysis of hydraulic fracture behavior with a
pseudo-three-dimensional simulator (in SPE production engineering 4 (1): 69 –74.
Nolte, K. 1979. Determination of fracture parameters from fracturing pressure decline. SPE 8341.
Proc., SPE Annual Technical Conference and Exhibition, Las Vegas.
Seth, Gaurav, Albert Reynolds, Jagannathan Mahadevan. Numerical Model for Interpretation of
Distributed-Temperature-Sensor Data During Hydraulic Fracturing.
Soliman, MY. 1986. Technique for Considering Fluid Compressibility and Temperature Changes in
Mini-Frac Analysis. Paper SPE 15370-MS presented at the SPE Annual Technical Conference and
Exhibition, New Orleans, Louisiana, 5– 8 October (Reprint).
Valkó, P., M.J. Economides. 1995. Hydraulic fracture mechanics, Wiley (Reprint).
Weng, X. 1992. Incorporation of 2D fluid flow into a pseudo-3D hydraulic fracturing simulator (in
SPE production engineering 7 (4): 331–337.
20 SPE-174858-MS

Appendix A– Input Data and Results of Simulation

Table 2—Input data for fracture propagation simulation of Cases 1 through 6


SPE-174858-MS 21

Table 3—Input data for thermal simulation of Case 6


22 SPE-174858-MS

Table 4 —Comparison of the results, taken from different propagation models

Table 5—Fracture propagation simulation results of Case 6

You might also like