Вы находитесь на странице: 1из 15

PO169539 DOI: 10.

2118/169539-PA Date: 8-May-15 Stage: Page: 84 Total Pages: 15

Diagnostic Fracture Injection Tests:


Common Mistakes, Misfires, and
Misdiagnoses
R.D. Barree, J.L. Miskimins, and J.V. Gilbert, Barree & Associates

Summary maximum rate allowed by available horsepower, or to 75% of the


Over the last 20 years, diagnostic fracture injection tests (DFITs) planned treatment rate of the main fracture, whichever is achieva-
have evolved into commonly used techniques that can provide ble. A constant rate is then held for 3 to 5 minutes (Step 2 in Fig.
valuable information about the reservoir, as well as hydraulic- 1), after which a rapid stepdown can be conducted. The stepdown
fracture-treatment parameters. Thousands of tests are pumped ev- test is separate from the actual DFIT, but it is commonly run at the
ery year in both conventional and unconventional reservoirs. end of the pump-in to determine perforation and near-wellbore
Unfortunately, many tests that are pumped provide poor or no frictional pressure losses. The rate is then immediately reduced to
results because of either problematic data acquisition or incorrect zero, the instantaneous shut-in pressure (ISIP, shown as Step 4 in
analysis of the acquired data. Fig. 1) is obtained, and the falloff pressure is monitored for as long
This paper discusses common issues and mistakes made while as possible or as long as necessary to acquire the desired data.
acquiring DFIT data. Guidelines on how to avoid these errors and Even when high friction is not expected, the stepdown is recom-
secure the best possible data are provided, including data resolu- mended to make identification of ISIP easier.
tion, pump rates, test duration, and fluid selection. Strategies are This paper is not intended to provide specific instructions on the
provided to estimate the time required to reach fracture closure basics of how to analyze DFIT tests because that information is
and establish stable reservoir transients for analysis. described in detail in other papers (Barree 1998; Craig and Brown
The last part of the paper addresses potential (and commonly 1999; Craig et al. 2005; Craig and Blasingame 2006; Barree et al.
observed) problems in the analysis of the DFIT. These issues can 2009). Rather, it aims to aid in the acquisition of useful and analyz-
be magnified in tight-gas and shale reservoirs because of the long able data and to provide hints to common problems that occur either
data-acquisition times and the subtle pressure transients that can during the pumping of the test or during the analysis of the results. It
occur. Specific issues that are discussed include poor instantane- is not uncommon to hear statements that disparage the test; however,
ous-shut-in-pressure data from perforation restriction, loss of most problems can be attributed to failed tests that do not provide
hydrostatic head, gas entry and the resulting phase segregation, any useful data because they were pumped or analyzed incorrectly.
the use of gelled fluids, and errors in after-closure analysis. Failure is a valid concern—these tests cost additional time and
money, and if they fail to provide usable information, that value is
lost or destroyed. However, frequently, it is the quality of the
Introduction acquired data and the manner in which the acquisition does (or does
Diagnostic fracture injection tests (DFITs) are small pump-in not) take place that are the root issues. If appropriate steps are taken
treatments performed to gather data to help design follow-up hy- and quality data are acquired, the tests are quite valid and correct
draulic-fracturing treatments, as well as to characterize the subject interpretation can shed light on what the reservoir is trying to say.
reservoir. DFITs have their basis in conventional mini-fracture This paper addresses these acquisition issues, provides guid-
treatments; however, they are subtly different and are intended to ance on acquiring quality data, and discusses some common mis-
acquire significantly more data for treatment design and reservoir takes that occur during analysis. If these steps are addressed, the
description. Conventional mini-fracture treatments have histori- success rate of the treatments should increase significantly. Most
cally been focused on acquiring treatment-design parameters, of the examples in this paper focus on unconventional, tight
such as fluid efficiencies and leakoff values; however, DFITs are hydrocarbon systems; however, all points are equally applicable
intended to expand that role and acquire additional data such as to conventional reservoirs.
reservoir pore pressure, closure and fracture gradients, process-
zone stresses, transmissibility values that can be converted into Recommended Acquisition Procedure
reservoir-permeability values, and leakoff mechanisms. The ar-
Every time fluid is pumped into a formation, there is a risk of for-
rival of unconventional reservoirs has added even more value to
mation damage. Therefore, fluid injections of any kind should be
DFITs because most of the information gained is comparable with
minimized and conducted under controlled conditions. However,
traditional pressure-transient tests, which are impractical to run in
at the same time, enough fluid must be injected under fracturing
tight-sand and shale systems because the time to analyzable pseu-
conditions to adequately contact the reservoir and provide condi-
doradial flow can be months, if not years.
tions under which the desired parameters can be measured.
A generic DFIT is shown in Fig. 1. The wellbore is first filled at
A diagnostic fracture injection test (DFIT) is a fracture test and
a low to moderate rate, until a positive surface-pressure response is
must be conducted at a stable fracturing rate. It does not take
observed. With low to moderate wellbore compressibility, the
extremely high rates to initiate and propagate a fracture in low- to
pressure should rise quickly until initial breakdown occurs. Break-
moderate-permeability formations. Even in permeabilities in the
down will be indicated either by a sharp drop in pressure as a new
millidarcy range, rates of 1 to 2 bbl/min are frequently high
fracture initiates (Point 1 in Fig. 1) or by a plateau in pressure as
enough to exceed fracturing pressure and cause fracture extension.
existing fractures are opened and extended. Once a breakdown
Fig. 2 shows an example relating the generated pressure for a vari-
event is observed, the injection rate should be increased to the
ety of reservoir permeabilities under a given set of reservoir condi-
tions. In an effort to pump enough fluid to adequately contact the
Copyright V
C 2015 Society of Petroleum Engineers reservoir and do so in a relatively short amount of time, rates of 5
This paper (SPE 169539) was accepted for presentation at the SPE Western North to 6 bbl/min (or even slightly higher at 7 to 8 bbl/min) have proved
American and Rocky Mountain Joint Regional Meeting, Denver, 16–18 April 2014, and to be very successful. This maximum rate is recommended to be
revised for publication. Original manuscript received for review 6 February 2014. Revised
manuscript received for review 6 September 2014. Paper peer approved 15 September
held for 3 to 5 minutes, as shown in Fig. 1, which results in approx-
2014. imately 600 to 1,800 gal of fluid being injected.

84 May 2015 SPE Production & Operations

ID: jaganm Time: 16:21 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 85 Total Pages: 15

1. Low rate to fill well and break down (t 0) 100


2. Hold constant maximum rate for 3–5 minutes
10
3. Step down to 75% then 50% of maximum
rate, 10–15 seconds for each step 1

Rate (bbl/min)
3 4. Shut in (for ISIP) and isolate gauge (t )
p
5. Record falloff as long as practical (t>t p) 0.1
0.01
4
2 0.001
1
5 0.0001
0.00001
0.000001
0.0001 0.001 0.01 0.1 1 10 100 1,000
Fig. 1—Generic DFIT procedure. Rate is plotted in black, with Permeability (md)
surface pressure in red. Initial breakdown is achieved at Point
1; a constant rate is held for 3 to 5 minutes, indicated by Point
2; Point 3 shows a rapid stepdown; Point 4 indicates ISIP; and Fig. 2—Rate necessary to initiate and extend a fracture at
Point 5 shows falloff. various reservoir permeabilities (absolute and undamaged).
Assumes a reservoir depth of 7,000 ft, fracture gradient of 0.8
psi/ft, normal hydrostatic pore gradient, formation thickness of
Data resolution for the entire injection and falloff period is 20 ft, 40-acre drainage area, and a water-injection system.
critical. Common pumping-gauge resolutions of 5 to 10 psi are
not adequate because the intent of the treatment is to look for
subtle variations in the derivative of the pressure vs. time. Low re- ity far-field reservoir zone (k ¼ 100 md), a near-fracture invaded
solution gauges can miss these data and render the test unusable. zone (k ¼ 1 md), and a thin-walled filter-cake zone (k ¼ 0.001
It is recommended to use gauges that have a resolution of at least md). In series flow, the total pressure drop through the system is
0.01 to 0.10 psi, with a sampling rate of one measurement per sec- the sum of the pressure drops through each zone (Willhite 1986).
ond during the pumping period and through fracture closure. This By use of Darcy’s law in a linear fashion, each pressure drop can
sampling rate can then be expanded during the falloff period after be determined from the length and permeability of each zone.
closure to one measurement every 30 seconds, which is adequate When even a thin film of very high flow resistance is present,
and achievable with appropriate gauge memory systems. such as the filter cake, the flow capacity of the least-conductive
Two key parameters in the analysis of a DFIT are the time of region dominates the system. When this filter cake is deposited on
injection and the volume of injection (Barree et al. 2009; Barree a fracture wall, as demonstrated in Fig. 4, most of the pressure
1998); therefore, the injection schedule must be recorded pre- drop is taken across the filter cake during leakoff. The far-field
cisely. Ideally, it is measured digitally in conjunction with the pressure gradient is much less than expected when computed on
pressure during the injection period. However, if this is not possi- the basis of the leakoff rate, and the after-closure analysis yields
ble or is not performed, extremely detailed notes must be kept an estimate of reservoir-flow capacity (i.e., kh) that is much too
that record the pumping start/stop times and any rate changes and high and is inconsistent with the observed closure time.
times of such changes made during the procedure. If the data are As a last acquisition component, isolation of the wellhead after
recorded, a rate schedule can then be recreated during analysis. pumping ends is required to eliminate any inadvertent pressure
Fluid selection for the treatment is a common question—what transients that might be caused by mechanical or logistical issues.
is the best option? In general, a Newtonian, nonwall-building fluid Because DFIT analysis uses derivatives, any large changes or
should be used. This is one area that differs significantly from tra- “bumps” in pressure can inadvertently alter the test and/or render
ditional mini-fracture treatments, which commonly inject gelled or it impossible to analyze. Pressure bleedoff that might occur during
other non-Newtonian fluids to measure fluid efficiency. For DFIT rigdown must be avoided. Likewise, if long-term gauges are being
purposes, water, diesel, or some other type of nonwall-building used that will remain in place after the pumping equipment is
fluid should be used. The reason for this is demonstrated by Figs. 3 released, these gauges must not be disturbed in any way.
and 4. Fig. 3 demonstrates the conventional fluid-loss model as a
1D solution for linear transient flow with constant-pressure bound- Common Acquisition Issues
ary conditions. The fluid pressure at the fracture face is assumed As noted, a diagnostic fracture injection test (DFIT) comprises
constant with time and the far-field pore pressure is assumed to be injecting water or other clean fluid at a pressure greater than the
constant. Initially, the pressure gradient and the associated leakoff fracture-extension pressure, creating a stable fracture geometry,
rate are very high. With time, the transient moves farther into the
reservoir and the gradient (and rate) decrease. This solution shows Pfrac
rate decreasing linearly with the square root of time.
ΔP3 = L3/k3 + ΔP2 = L2/k2 + ΔP1 = L1/k1 = ΔPT
The inset of Fig. 4 then shows leakoff modeled as a combina-
tion of series flow. The figure roughly describes a high-permeabil- 10 2 0.5

k = 100 k=1 k = 0.01


Pfrac
kavg = Lt/L1/k1+L2/k2+L3/k3) = 0.025
Early
time Later
time
Ppore
Distance from fracture face

Fig. 4—When a filter cake is deposited on a fracture wall (left of


the figure in blue), most of the pressure drop (shown as the red
Ppore line) is taken across the filter cake during leakoff. This is unlike
Distance from fracture face the behavior shown in the inset, which demonstrates the total
steady-state pressure drop in an undamaged system (DPT),
Fig. 3—Conventional fluid-loss model as a 1D solution for lin- modeled as a combination of series flows for permeabilities of
ear transient flow with constant-pressure boundary conditions. 100 md (k1), 1 md (k2), and 0.001 md (k3) (Willhite 1986).

May 2015 SPE Production & Operations 85

ID: jaganm Time: 16:21 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 86 Total Pages: 15

5,500 when the treatment is pumped. Because these secondary gauges


are calibrated differently than the pump gauges, if the two data
5,000 sets are overlapped, there will often be a difference in the read-
ings. Any difference in readings will show as a slope change, will
4,500 affect the pressure derivatives, and will provide false indications
Treating Pressure (psi)

of reservoir conditions during analysis. The secondary long-term


4,000 gauges must be opened to the wellhead during the pumping pe-
riod, to eliminate this issue.
3,500 Additional field operational issues that can cause acquisition
problems are frozen or temperature-affected wellheads and offset
3,000 reservoir interruptions. Precautions should be taken for wells in
winter conditions to prevent freezing of the surface gauges. Even
2,500 if the gauges do not freeze completely, temperature fluctuations
can impact the pressure readings, and insulating wellheads is of-
2,000 ten helpful to prevent these pressure movements. Thermal fluid
0 5 10 15 20 25 30 35 40 expansion or contraction in the wellbore will affect both surface
Pump Rate (bbl/min) and downhole gauges. Some assume that a downhole gauge will
Step Up Step Down be in a constant-temperature environment, and will therefore not
be affected. In a low-compressibility-fluid-filled wellbore, the
Fig. 5—A step-rate analysis chart showing increasing rate coefficient of thermal expansion, coupled with density-driven
(stepup) measurements vs. decreasing rate (stepdown) meas-
convection currents, can still cause a pressure change throughout
urements. Note the hysteresis exhibited between the two
curves. the wellbore volume (Tompkins et al. 2014).
Again, the behavior of pressure transients is critical to that
analysis of the treatment—if the pressure changes are caused by
shutting down injection and studying the pressure decay to evalu- anything other than the reservoir and fracture response, the treat-
ate leakoff mechanism and find fracture closure, and then evaluat- ment can provide false readings. This is also the case if the reser-
ing the after-closure pressure decline with appropriate reservoir voir itself is upset by any pressure changes such as pumping
transient solutions. If any of these steps is compromised during offset hydraulic-fracturing treatments. A quiet environment in
acquisition of the data, the data can be unusable. both the surface and downhole environments is needed for opti-
As demonstrated in Figs. 3 and 4, a Newtonian, nonwall-build- mal data acquisition.
ing fluid must be used when pumping the treatment, if any reser- One final acquisition precaution relates to the stepdown por-
voir properties are to be obtained from the analysis. The use of tion of the test. It is strongly recommended to use a stepdown test
gelled or other non-Newtonian fluids can disrupt the after-closure vs. a stepup test for the reasons displayed in Fig. 5. Fig. 5 shows
pressure gradient and mask the reservoir flow capacity. Another the treating pressure for increasing rates (overlapped by the pink
common issue that is also a holdover from old-style mini-fracture line) and then the treating pressure for decreasing rates. The hys-
treatments is the use of multiple pump-ins. DFITs, especially in teresis in the curve is obvious and indicates that during the initial
low-permeability systems, should be considered as “one and rate-increase period, the wellbore is still experiencing breakdown
done.” If the treatment is not pumped correctly the first time, an episodes and changes in the overall net pressure and fracture-
immediate second attempt is almost pointless because the reservoir extension period. In many cases, the number of perforations open
conditions have now been altered by the first injection and the cre- to injection changes with each increase in pump rate. The step-
ated pressure transients will not have dissipated by the time the down period does a better job of measuring the frictional-change
second attempt is usually made. The assumptions required for information that is desired rather than the complexity of friction
valid analysis of a pulse test do not allow for superposition of mul- and fracture-extension components combined.
tiple injections. It is possible that a second attempt can be made if
the first attempt was not too invasive, but one must remember that Common Analysis Mistakes
it may take days, if not weeks, for the pressure transient to fully
dissipate and for the reservoir to reach unperturbed conditions. In addition to the acquisition issues mentioned previously, there
Strictly adhering to the assumption of a single pulse injection are also several common analysis mistakes after the diagnostic-
into a constant potential reservoir system may be impossible, but fracture-injection-test (DFIT) data are acquired. The following
all reasonable attempts should be made to minimize perturbations subsections discuss some of these more-common errors and ways
to the system. Pressure transients may persist after perforating or that they can be avoided (if possible).
perforation breakdown. A serious upset to the system can be
induced when pressure-actuated sliding sleeves are used to test Incorrect Instantaneous-Shut-in-Pressure (ISIP) Determination.
the toe of a horizontal well. Wellbore-fluid decompression when Analysis of a pressure falloff begins with the determination of the
the sleeve opens can induce a large injection volume, pressure ISIP. This value is taken to represent the fracture-extension pres-
transient, and fracture extension before the beginning of the sure, when corrected to bottomhole conditions, and the fracture-
DFIT. If possible, the DFIT should either be combined with the extension gradient (Barree et al. 2009). The difference between
initial perforation-breakdown operations or be delayed long the ISIP and the determined closure pressure is taken as the net
enough after perforating to allow the induced pressure transient to pressure in the fracture, or as a measure of the resistance of the
stabilize. Similarly, some slight pressure transients may be in- rock to the facture extension or process-zone stress (PZS).
duced if the hole needs to be filled before the test. In most cases, Because net pressure determines fracture width, and to a large
this operation will be (or should be) conducted below fracturing degree height (because of the expectation that confining stress
pressure so that injection into the reservoir is minimized. contrast controls containment), the accurate determination of ISIP
Not acquiring falloff data for a long enough period is also a is critical to any later analysis and fracture-geometry modeling.
common acquisition issue. How long to plan on letting the falloff In many cases, the ISIP is not obvious and does not occur im-
period be recorded for different reservoir types is discussed in the mediately after shut-in. When a substantial near-well pressure
following sections of this paper; however, it suffices to say that drop exists, either through poor perforation efficiency or high
the time period will be longer than the amount of time the pump- near-wellbore tortuosity, injection at the fracture entrance may
ing equipment will physically stay on location. Therefore, second- not stop for some time after the surface injection stops. The
ary gauges that remain after the pumping equipment is removed amount of continued injection after shut-in, the amount of excess
are generally required. One issue with these secondary, long-term frictional pressure that must dissipate, and the time required for
gauges is that they are occasionally not opened to the wellhead decompression of the wellbore fluids depend on the severity of

86 May 2015 SPE Production & Operations

ID: jaganm Time: 16:21 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 87 Total Pages: 15

Pressure/Rate Events
Pressure (psi) Time Pressure Rate
Rate (bbl/min) Test Events ISIP 10.233 0.164

1700 2050 2400 2750 3100 3450 3800 4150 4500


ISIP

4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0
Pressure (psi)

Rate (bbl/min)
6.00 8.00 10.00 12.00 14.00 16.00 18.00 20.00 22.00 24.00
Time (minutes)

Fig. 6—Rate and pressure chart for a test with extremely high entry friction. (No ISIP value is noted in the upper right corner
because the ISIP is not yet picked on the figure.)

the near-well restriction to flow and the compressibility of the The thin blue line in Fig. 7 is the measured surface pressure
wellbore volume. for the test, and the heavy line is the computed wellbore pressure
Fig. 6 shows a fairly extreme case with an apparent ISIP of with the frictional pressure drop calculated from a constant tortu-
approximately 3,640 psi at the surface. During injection, there is a osity factor. The injection rate declines from 4 bbl/min at the
breakdown, followed by fluid decompression, and then repressuri- instant of shut-in to 1.0 bbl/min after 25 seconds, and finally to
zation of the well. The cause of the repressurization is unknown, less than 0.05 bbl/min after 200 seconds. Perforation pressure
but is probably related to mechanical perforation restriction. The drop and other factors are ignored in this analysis. The main
two pressure-decay trends are parallel, indicating the same well- point is that wellbore storage and decompression can drive injec-
bore-dominated process controls the response. The second and tion after shut-in for a long time (several minutes) and can
final pressure falloff has been analyzed. The pressure falls nearly account for large pressure drops. With large wellbore volumes,
1,800 psi in the first few minutes after shut-in, with an injection typical of extended-reach horizontal wells and larger pipe sizes,
rate of only 4 bbl/min. This DFIT was conducted through perfora- this effect becomes more pronounced. Any gas or more-com-
tions in the toe of a cemented horizontal well. The inside diameter pressible fluid in the wellbore will extend the blowdown period
of the pipe was 4 in., and the total depth of the well was approxi- even longer.
mately 15,000 ft. The wellbore was filled with fresh water for the Now that the source and duration of the early pressure decay is
test. Before any meaningful analysis could be performed, an ISIP accounted for, what is the actual representative ISIP? One method
had to be selected somewhere between 3,640 and 1,750 psi. If a that can help is to extrapolate the pressure-vs.-log-of-time curve,
value significantly below the pressure at shut-in (3,640 psi) was as shown in Fig. 8, back to the point of shut-in. This construction
used, the excess pressure drop had to be explained. suggests that some effect of wellbore-fluid expansion may be felt
The total compressibility of the wellbore described can be esti- for as long as 18 minutes after shut-in, and gives an estimated
mated from the water compressibility (3.0106 psi) and well- effective ISIP of 1,766 psi. This may be a lower-bound value of
bore volume (233 bbl). Expansion of the pipe should also be ISIP, but it is a good starting estimate.
considered. By use of the thin-walled cylinder equation (Eq. 1; The validity of the ISIP estimate can often be verified by ex-
Young and Budynas 2011), the change in volume of the wellbore amination of the log-log plot of pressure change from the assumed
can be estimated, where d is the pipe diameter (in.), Lthickness is ISIP after shut-in and the semi-log derivative of the pressure-dif-
the wall thickness (in.), Pinternal is the internal pressure change ference curve vs. shut-in time. This plot, shown in Fig. 9, should
(psi), and Esteel is the Young’s modulus of steel (30.0106 psi): produce parallel lines of the pressure difference (blue DP) and de-
rivative (red) during the preclosure period. The magnitude of the
Pinternal d2 derivative should be the pressure difference multiplied by the
¼ Dd: . . . . . . . . . . . . . . . . . . . . . . . . ð1Þ
2Lthickness Esteel slope of the derivative curve (Barree et al. 2009). In the example,
the pressure-difference curve is 300 psi at closure and the deriva-
Combining the water and pipe compressibility, the pressure of tive is roughly 0.28 (the slope) times that value. Note that any
the fluid in the wellbore will change by approximately 1,170 psi change to the estimated ISIP value will move the pressure-differ-
for every 1 bbl of volume change. This relationship can be used ence curve, but will have no effect on the derivative-curve slope
with an estimate of pressure drop because of tortuosity, assuming or position on the plot. When a stable linear derivative is apparent
a square-root-of-rate linear relation, to estimate the frictional pres- through the preclosure data, this method provides an effective
sure drop driven by fluid expansion, and the expansion rate after final check, but is not infallible.
shutdown. With an injection rate of 4 bbl/min at shut-in, a fric-
tional pressure drop of 1,940 psi results from a tortuosity factor of
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Basic Selection of Closure. Basic DFIT analysis is covered in
970 psi= ðbbl=minÞ. In the first second after shut-in, this rate detail in Barree et al. (2009), but some mistakes in these techni-
decreases the wellbore volume by 0.0667 bbl, causing the pres- ques continue to be made by many analysts. A general comment
sure to drop from 3,640 to 3,562 psi. As the fluid continues to regarding fracture “closure” must be made because it affects the
expand, the injection rate drops, and the inlet frictional pressure interpretation of the test results. Because rocks are not perfectly
decays. Fig. 7 shows a plot of the pressure decay computed from elastic, there will always be some residual deformation after open-
this simple relationship and compared with the observed pressure ing a hydraulic fracture. The initial pressure decline, before clo-
decline in the test. sure, is dominated by the rebound of the compressed rock mass

May 2015 SPE Production & Operations 87

ID: jaganm Time: 16:21 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 88 Total Pages: 15

Diagnostics

1600 2000 2400 2800 3200 3600


Gauge_Pres

4.5
Pump Rate
Derived WH_Pres
Derived BH_Pres
Average Rate

4.0
3.5
Rates (bbl/min)
3.0
Pressure (psi)

2.5 2.0
800 1200

1.5
1.0
400

0.5
0.0
0.41 0.42 0.43 0.44 0.45 0.46 0.47 0.48 0.49 0.50
Elapsed Time (hours)

Fig. 7—Actual (thin blue line) and computed (thick blue line) pressure decay after shut-in, accounting for near-well tortuosity and
wellbore decompression.

surrounding the fracture. As the rock returns to its initial, unde- Fig. 10 shows a DFIT with variable fracture compliance (or
formed state, the strain energy related to net stress decreases with storage during closure), where the closure time and pressure have
fracture aperture. Once the rock has fully relaxed, it stops actively been mistakenly diagnosed. In the original analysis, the closure
displacing fluid from the fracture and the rate of pressure decay time and pressure were identified by the positions of the red lines,
changes because reservoir transients begin to dominate the falloff. at a closure time of G ¼ 14 and P ¼ 3,960 psi; however, the cor-
This point, where the internal fluid pressure balances the Earth rect closure time is G ¼ 32 and P ¼ 3,640 psi (the process for
stress (zero net stress on the fracture face), is what is called chosing the correct closure is discussed in detail in Barree et al.
“closure,” but the fracture is not closed or sealed mechanically. 2009). This discrepancy will obviously affect any interpretation
The entire face of the fracture remains open to transmit the pres- of closure stress for calibration of a geomechanical model or
sure transient to the surrounding reservoir. Fluid movement along stress profile, and will impact estimates of formation flow
the length of the fracture and related pressure transients in the capacity and post-fracture production. The red arrows in Fig. 10
fracture are negligible, except in the case of continued fracture-tip indicate other incorrect closure events that are often selected by
extension after shut-in. The change in the pressure/time derivative mistake. The one correct closure event is shown by the blue
response corresponding to the loss of energy input from the arrow, after the end of the variable compliance and linear closure
deformed rock mass around the fracture is detectable with all periods. The error in closure stress also leads to an error in the
DFIT analysis techniques. interpretation of fracture net-extension pressure. This affects the

Test Events
Pressure/Rate Events
Pressure (psi) Time Pressure Rate
Rate (bbl/min) ISIP 10.233 1766.86 0.164
1700 2050 2400 2750 3100 3450 3800 4150 4500

ISIP
4.5
4.0
3.5
3.0 2.5 2.0 1.5
Pressure (psi)

Rate (bbl/min)
1.0

ISIP = 1766.86
0.5
0.0
00

00
00

8. 0
00

70 0
0

10 00
00

0
0

.0

.0

.0

.0

.0

.0
.0

0
4.

5.
6.
7.

.
0.

0.
10

20

30

40

50

60

90

20

Time (minutes)

Fig. 8—Extrapolation of the pressure-vs.-log-of-time to approximate ISIP.

88 May 2015 SPE Production & Operations

ID: jaganm Time: 16:21 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 89 Total Pages: 15

Log-Log Analysis Closure Events


dT BHP DP
C Closure 422.949 5903.35 336.626
ΔP (psi)
DT dDP/dDT BL Begin Linear Flow 1132.294 5813.58 426.396
EL End Linear Flow 2853.175 5755.66 484.316

600
c BL EL
500
400
300

200

100 0
Pressure (psi) 0.28
–0.
500

10

10.00 100.00 1000.00


dT (minutes)

Fig. 9—Log-log plot of pressure difference and derivative vs. time. The pressure difference (blue curve) is roughly 300 psi at clo-
sure (C), and the derivative value (red curve) should be this value times the slope (0.28) (i.e., 84).

predicted width of the fracture, and therefore directly impacts the that the net pressure must be higher for a given fracture width. A
computed fracture length through conservation of injected vol- higher net pressure implies the ability to break through more bar-
ume. Predicted fracture height may also be affected, depending on riers and possibly generate more fracture height, thus increasing
the strength and stress of the confining layers. In short, if the compliance. The system should stabilize at an equilibrium frac-
wrong analysis results are applied to the calibration of any frac- ture height, width, and net pressure for a given injection rate and
ture design model, the results will be incorrect. fluid system.
For a DFIT to be valid, an injection rate must be selected that
will generate a net pressure and fracture geometry that is repre-
Effect of Injection Time vs. Pump Rate. Assuming a high sentative of the main fracture. In practical terms, this means that
enough injection rate is used so that the rock is dilated to its ulti- the highest possible injection rate for the test will lead to the most
mate capacity, the width of the fracture will be determined by the representative results. Once a stable net pressure and fracture ge-
net pressure generated, modulus of the rock, and some character- ometry are achieved, a higher injection rate will cause the fracture
istic dimension of the fracture. The theoretical basis of the G- to grow faster, but should not alter the width or height signifi-
function analysis assumes that the fracture adheres to Perkins- cantly. This assumption deviates from the classical PKN assump-
Kern-Nordgren (PKN) geometry assumptions (Nolte 1979). While tions, wherein all net pressure is generated solely through the
this assumption is not necessarily valid, it implies that the gener- frictional pressure gradient of the moving fluid. More-realistic
ated fracture height will be the controlling dimension. This leads analysis of DFIT results requires that there be some resistance to
to a somewhat circular argument because net pressure, which is fracture propagation inherent in the rock. This has been referred
proportional to fracture width, and fracture compliance (H/E or to as PZS, “fracture toughness” and “rock strength,” and by other
fracture height/Young’s modulus) are directly linked in this names, although each carries certain different connotations
model. A confined height (H) for a constant modulus (E) means regarding the mechanism of resistance to fracture growth. A very
low injection rate will develop less net pressure, therefore less
fracture height and width, and will not strain the rock to its ulti-
mate limit. This can result in missing the evolution of secondary
3200 3400 3600 3800 4000 4200 4400

700

Not Closure shear fractures that can affect compliance, leakoff rate, and sys-
Correct Closure tem permeability. The original theoretical models assumed the
600

rock to be linearly elastic and homogeneous, with no considera-


tion that shear failure could exist.
400 500

Assuming a stable net pressure and fracture geometry are


GdP/dG

achieved, the fracture volume to surface area (or fracture width)


will remain constant for any duration of injection, which implies
that a fracture with a height of 50 ft and a length of 300 ft will
300

close in the same time as a fracture that is 50 ft tall and 600 ft in


length, as long as leakoff flux per unit area is the same in both
200

cases. For this reason, the time to closure is not a function of


injected volume, and is insensitive to rate as long as a high
100

enough rate is achieved to reach a stable failure and fracture-prop-


agation condition. In fact, time to closure is a function of pumping
10 20 30 time and not volume injected.
G (dt) The theoretical calculation for time to closure depends on the
integration of the leakoff rate through each element of the fracture
Fig. 10—G-function derivative plot for variable compliance dur- face as it is being created and during the closure process. For sin-
ing closure, with incorrect closure picks indicated by red gle-phase flow with no damage to the face of the fracture (such as
arrows and the correct closure pick shown by the blue arrow. a gel filter cake), the leakoff rate through each element of the

May 2015 SPE Production & Operations 89

ID: jaganm Time: 16:21 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 90 Total Pages: 15

3000
1
3
2500 5

Fracture Half-Length (ft)

Pump Rate (bbl/min)


8
2000 10
12

1500

1000

500

0
0 2 4 6 8 10 12 14 16 18 20
Case A Pumping Time (minutes)

Fig. 12—Fracture half-length created as a function of pumping


time and rate. Assumes a constant fracture height of 30 ft, a
fracture width of 0.1 in., and 100% fluid efficiency.

diagnostics. These tests seek to approach a matrix injection/falloff


Case B test and should not be considered as valid DFITs because the in-
formation critical to fracture design will be compromised.
Fig. 11—Pressure transients established by low-rate injection, Another failing of the low-rate test is that the height of the zone
Case A (upper figure), and high-rate injection, Case B (lower contacted will rarely (in the real world that means, effectively,
figure). never) be the perforated pay height. Therefore, any permeability
or even kh estimate will probably not match post-fracture
production.
fracture face should decrease linearly with the square root of time
since exposure of that face element. For a slow injection rate,
each element of the face of the fracture is exposed for a longer Size of Generated Fractures. Most people consider a short-
time, so the leakoff rate from the total fracture face will be much duration fracture injection test to be too small to effectively evalu-
slower at shut-in than for the high-rate-injection case. ate enough reservoir volume to be applicable for post-fracture
Fig. 11 shows conceptually the radius of investigation of the production. Fig. 12 shows the estimated fracture half-length that
leakoff pressure transients at three times for a low-rate-injection would be generated for a 30-ft fracture height, with an average
scenario (Case A) and a high-rate-injection scenario (Case B). fracture width of 0.1 in., given 100% fluid efficiency. Fracture
The centerpoint represents the wellbore or point of fracture initia- half-length is then L¼qitp/(2HW) in consistent units. These may
tion. The inner set of black points represents the fracture half- seem to be arbitrary input assumptions, but they are reasonable
length for some point near the middle of the injection period. In for water injection in “typical” rocks with Young’s modulus of 3
Case A, the pressure transient has penetrated a much larger dis- to 4 million psi and with a net pressure of 1,000 psi and very low
tance into the formation because of the longer injection time at system permeability. The estimated fracture length can be directly
the low rate, and the leakoff rate is much lower than in Case B. In multiplied by fluid efficiency to correct for leakoff in higher-per-
Case B, the same fracture length is reached at a much shorter meability systems, and corrected by the ratio of created to
injection time, so the pressure transient cannot travel as far from assumed fracture height, if it is known.
the fracture face, and the leakoff rate is faster than in Case A. The Note that for the recommended rate of 10 bbl/min, a pumping
outer set of black points represents the fracture position at the end time of 5 minutes generates a fracture half-length of more than
of injection. In Case A, the pressure transients along the length of 600 ft, when leakoff during the 5-minute injection period is negli-
the fracture have traveled a greater distance into the formation, gible. This is very likely the case in shale or ultratight sand or car-
leakoff rate through each element of the fracture surface is much bonate reservoirs. Extending the pumping time to 10 minutes
lower (following the square-root-of-time function), and therefore increases the created length to almost 1,200 ft and increases the
time to reach closure is significantly longer than in Case B. In the time to closure by increasing the average exposure time of all
high-rate case, the entire fracture length is established with a short fracture-face elements. Clearly, in tight rocks, these are not
time of exposure of all elements of the fracture surface, and the “small” fractures, and the formation surface area exposed is statis-
leakoff rate through each element remains high. tically significant. For the 5-minute, 10-bbl/min injection case, the
For the same injected-fluid volume, the fracture in Case B may surface area of one of the four fracture-wing faces exposed is
be slightly longer than the Case-A fracture, but the difference is 18,000 ft2.
not linear with rate. In Case B, the leakoff rate is higher through-
out the injection period, but the fluid efficiency is also higher. All
this assumes a similar net pressure and fracture compliance in Design of Test Duration. Longer pumping times allow more
both cases. time for penetration of the induced pressure transient. Given a sta-
A second result of the impact of rate is apparent from Fig. 11. ble fracture geometry, the time required to close the fracture by
In the low-rate (Case A) condition, the pressure transient at clo- leaking off a fixed volume/surface-area ratio of fluid in the frac-
sure, represented by the outer ellipse, has reached a nearly pseu- ture can be related to the system permeability. A good empirical
doradial-flow case very quickly after closure. In Case B, the approximation for time to closure that is based on literally thou-
transient remains in pseudolinear flow for a very long time after sands of analyses is given by
closure. For this reason, many people (generally those concerned tp
with application of classical pressure-transient theory) recom- tc ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð2Þ
3k
mend very low injection rates. This recommendation is based
solely on the desire to establish a pseudoradial-flow period within where tc is the time to reach the fracture closure (minutes); tp is
a practical test duration. Unfortunately, in the real world (where the pumping time (minutes); and k is the permeability of the leak-
rocks are not isotropic, homogeneous, and linearly elastic and off system connected to the fracture face, including any secondary
have no inherent resistance to fracture growth), the use of a very permeability enhancement resulting from induced shear failure of
low injection rate makes the test essentially useless for fracture the host rock. Fig. 13 is a plot of the estimated closure time for

90 May 2015 SPE Production & Operations

ID: jaganm Time: 16:22 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 91 Total Pages: 15

1000.00 2 the closure time has elapsed. So, with the 0.01-md, 5-minute-
5 injection case, allow 2 to 3 hours for closure and 6 to 9 hours after
10
20 shut-in to the start of a valid reservoir-flow regime. For the 0.001-
100.00
Hours To Reach Closure
md case, that is 24 hours to closure and 3 days of shut-in for any

Pump Time (minutes)


useful reservoir information.
10.00 A good example of the time lag required to reach a stable res-
ervoir transient is shown in Fig. 14. After some nonlinear-leakoff
behavior (including some mechanical noise in the data), there is
1.00 an obvious break in the semilog derivative curve at the closure
time of 52 minutes after shut-in. The derivative continues to curve
0.10 until the first indication of a valid reservoir pseudolinear-flow
transient at approximately 185 minutes after shut-in. The deriva-
tive continues to steepen, and approaches a negative unit slope
0.01 (not shown on the plot because the derivative has not reached a
0.0001 0.001 0.01 0.1 1
System Permeability (md)
true unit slope) that some may be tempted to analyze as a pseu-
doradial reservoir-flow regime. For the purposes of this discus-
Fig. 13—Approximate time to reach fracture closure for various sion, the important consideration is that any reservoir transient
pumping times for a range of reservoir-system permeabilities. must take time to develop. Expecting a sharp break in the deriva-
tive curve immediately after closure to represent a reservoir-flow
transient is simply not reasonable.
pumping times of 2 to 20 minutes over a range of system perme- If we follow the assumptions for the PKN-fracture geometry
abilities from 100 nd (0.0001 md) to 1 md. For convenience, the and assume that linear-transient flow governs fracture-leakoff
time scale has been converted to hours of shut-in. flow up until closure, then the time required for the induced pres-
In these moderately high-permeability formations, where this sure transient to establish a pseudoradial-flow regime can be com-
technique was first applied, extending the pumping time may not puted from the definition of tDxf. For pseudoradial flow, a value of
be a serious concern. For this 0.1-md rock, a pump time of 20 tDxf ¼ 1.0 (at least) must be reached in the falloff period. Eq. 3
minutes will still reach fracture closure in approximately 1 hour (Cinco-Ley and Samaniego V. 1981) gives the time, in hours, to
of shut-in time. Note that for higher-permeability formations, clo- reach pseudoradial flow (tpr) as a function of reservoir storativity,
sure can happen too quickly to be easily identified when other fac- permeability, and fracture length:
tors, such as wellbore friction and compressibility, are considered.
At 0.01 md, the 5-minute pumping time requires 2 to 3 hours /lCt x2f tDxf
to reach closure. For a 0.001-md formation, which is not atypical tpr ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . ð3Þ
0:0002637k
for what are considered commercial tight silt or shale reservoirs,
closure can take up to 24 hours after a 5-minute injection. If the where tpr is the time to reach pseudoradial flow (hours), / is the
system permeability is actually in the 100-nd range, closure can porosity (fraction), l is the fluid viscosity (cp), Ct is the total reser-
take up to 10 days following a 5-minute injection. When system voir compressibility (1/psi), xf is the fracture half-length (ft), and k
permeability is this low, the created fracture length will be large, is the permeability of the leakoff system (md). Eq. 3 has been
and the pressure transient established at closure will remain in a solved for tDxf ¼ 1 for a gas-reservoir case with viscosity of 0.02
pseudolinear-flow regime for a very long time. cp, porosity of 8%, and total reservoir compressibility of
Note that for a 5-minute injection period into a 0.1-md forma- 3.0105 psi. The time, in days, required to reach the fully devel-
tion, fracture closure occurs in approximately 0.2 hours (12 oped pseudoradial-flow period according to these assumptions is
minutes). Once the fracture has closed, some time is required for shown in Fig. 15. Daltaban and Wall (1998) suggest that pseudora-
a stable reservoir transient flow regime to be established. On a dial flow may not be identified until tDxf ¼ 2 for an infinite-conduc-
log-log plot of pressure change vs. shut-in time, a safe and realis- tivity fracture or tDxf ¼ 3 for a uniform-flux fracture. The
tic lag time is one-half of a log cycle in time after closure, which assumptions used in Fig. 15 are, therefore, somewhat optimistic.
means (roughly) that no data should be analyzed for any reser- Consider a moderate permeability DFIT in a 0.1-md system
voir-flow regime, including pseudolinear flow, until three times with a pump time of 5 minutes at 10 bbl/min. The generated

Log-Log Analysis Closure Events


dT BHP DP
C Closure 52.174 5659.17 485.664
ΔP (psi)
BL Begin Linear Flow 184.736 5326.80 818.035
DT* dDP/dDT
EL End Linear Flow 394.399 5174.65 970.186
BR Begin Radial Flow 1694.648 5024.62 1120.21
2000

C BL EL
1000

–0.
500
Pressure (psi)

0
0.50
100

1.00 10.00 100.00 1000.00


dt (minutes)

Fig. 14—Log-log pressure change vs. shut-in time plot showing closure (C) and start of pseudolinear flow (BL).

May 2015 SPE Production & Operations 91

ID: jaganm Time: 16:22 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 92 Total Pages: 15

100,000
0.0001
10,000 0.001
0.01

System Permeability (md)


1,000 0.1
Days to Radial Flow

1
100

10

0.1

0.01

0.001
10 100 1,000 Fig. 16—Complex, nonplanar fracture system (black lines) with
Fracture Half-Length (ft) coalescing transient pressure fields (blue ellipses).

Fig. 15—Time, in days, to theoretically reach pseudoradial res-


ervoir transient flow conditions for a gas reservoir for various forations or to achieve uniform injection over a predictable forma-
system permeabilities. tion height.
Why do some DFITs appear to enter a pseudoradial-flow pe-
riod based on the negative unit slope of the semilog derivative on
fracture will close in approximately 15 minutes and have a fluid
the log-log pressure/time plot or based on the after-closure flow-
efficiency of approximately 50%. According to Fig. 12, the gener-
regime-analysis plot? The most plausible answer is that fractures
ated fracture half-length would be approximately 300 ft for a frac-
do not adhere to the simple planar geometry of a PKN fracture
ture height of 30 ft (the default assumed height in the plot). By
that is assumed in the derivation of the G-function equations and
use of the rule of thumb for start of after-closure analysis, a pseu-
the conventional after-closure solution of the diffusivity equation.
dolinear transient may develop approximately 45 minutes after
A more likely or at least plausible image of the fracture system
shut-in. If the created fracture is planar, constant height, and fol-
created in an unconventional reservoir is shown in a map view in
lows the assumptions for a homogeneous system permeability
Fig. 16.
under linear flow during closure, Fig. 15 shows that a pseudora-
If a complex, nonplanar fracture system develops during injec-
dial-flow regime should theoretically not be established until 2 to
tion, leakoff will occur through each face of all fracture strands.
3 days after shut-in.
Even with high volumetric fluid efficiency, the primary fracture
Conducting the same analysis with a 0.001-md system, this
will be much shorter than predicted for a planar fracture, as shown
permeability leads to a predicted closure time of 27 hours, start of
in Fig. 11, and normally assumed in theory. Time to closure,
pseudolinear flow after 83 hours (3.5 days), and start of pseudora-
based on the G-function, will be proportional to the overall sur-
dial flow after approximately 1,000 days of shut-in for a created
face-area/fracture-volume ratio at shut-in, whether the fracture is
fracture half-length of approximately 500 ft (83% efficiency).
planar or not. For this reason, the permeability estimated from the
This implies that expecting pseudoradial-flow behavior in a DFIT
G-function closure time and given in Barree et al. (2009) is a reli-
in unconventional reservoirs is unreasonable. No combination of
able estimate of the permeability of the system contacted by the
pump rate and pumping time will establish a stable fracture geom-
induced fracture(s). The leakoff rate will be affected slightly by
etry and result in stabilized pseudoradial flow in a practical test
the overlapping of the transient-pressure envelopes around each
duration with system permeability less than approximately 0.01
fracture, but the time required to leak off the fluid volume in the
md. Attempting to induce early onset of pseudoradial flow by con-
fracture system should be related to net pressure and system-flow
ducting the test at very low rate or very short duration compro-
capacity or permeability. Some crossflow between fracture strands
mises the geometry of the fracture and the validity of any data
opening against different normal stresses should be expected.
associated with fracture-extension criteria. If a test is to be con-
This can lead to the variable storage signature seen in G-function
ducted as an injection/falloff test for reservoir properties, then it
derivative analysis. Low compliance in the fracture systems
should not be confused as a diagnostic of fracture-extension or -
coupled with a large, exposed surface area can also lead to the
closure conditions. Low-rate and small-volume tests are then
pressure-dependent permeability signature. Because both of these
compromised by the inability to achieve breakdown of many per-
phenomena are commonly observed, it is reasonable to assume
that complex fracture systems in unconventional reservoirs are
common.
zNRT As shut-in time progresses, the overlapping transient-pressure
P2 = P1 =
V envelopes will coalesce into what appears to be a more-radial-
flow geometry, illustrated by the outer ellipse in Fig. 16. As this
pressure transient dissipates into the far field, it will appear as a
radial-flow signature on the diagnostic derivative plots. However,
this radial-flow pattern does not conform to the assumptions in-
herent in the after-closure pseudoradial-flow solution. This solu-
Phead = 0.45 psi/ft tion is based on the assumption of a single planar fracture with a
transient evolving from a linear-flow pattern along the face of a
single created fracture, which transitions to a radial-flow pattern
over a dimensionless time proportional to the assumptions in Eq.
3 and Fig. 11. The net result is that an apparent pseudoradial-flow
regime develops much faster than would be the case for a single
zNRT planar fracture of equivalent fluid efficiency. Any transmissibility
P2 = Phead = derived from the apparent pseudoradial-flow time function may
V be orders of magnitude too high to represent the actual flow
capacity of the reservoir system. The problem and errors in inter-
Fig. 17—Schematic showing that a gas bubble that enters from pretation are compounded when the radial-flow or Horner-analy-
the perforations can generate increased pressures as it rises to sis techniques are applied when a true pseudoradial-flow pattern
the surface. does not exist at all.

92 May 2015 SPE Production & Operations

ID: jaganm Time: 16:22 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 93 Total Pages: 15

A Wellhead Pressure (psi) A Slurry Rate (bbl/min) B B


6000 3.5

3.0
5000

2.5
4000

2.0

3000

1.5

2000
1.0

1000
0.5

0 0.0
22:00 00:00 02:00 04:00 06:00 08:00 10:00 12:00 14:00 16:00 18:00
3/22/2003 3/23/2003 Time 3/23/2003

Fig. 18—The complete data set for a long-term injection test in an overpressured formation is shown. The minimum pressure
recorded at the surface was 743 psi after 2.5 hours of falloff. The pressure increases after that time until approximately 9 hours of
falloff. For purposes of analysis, the end of the valid data occurs before the minimum pressure point at 2.5 hours. The later pres-
sure rise is caused by entry of small gas bubbles at the perforations, which rise in the wellbore fluid column.

Mini-fracture Events
Time BGP SR
1 Start 10/2/2006 19:19:03 6393 6.100
2 Shut In 10/2/2006 19:35:34 4368 0.000
BH Gauge Pressure (psi) A 3 Stop 10/2/2006 23:23:39 3027 0.000 B
A Slurry Rate (bbl/min) B
6500 7
1 3
6000 2 6
5500
5
5000
4
4500 (ISIP = 4367)
3
4000
2
3500

3000 1

2500 0
10/3/2006 10/4/2006 10/5/2006 10/6/2006 10/7/2006
Time

Fig. 19—Bottomhole-gauge response with falling fluid level or surface pressure in vacuum.

2
Bottomhole ISIP - 4352 psi 1

1,000 9 (168.5, 1099)


8
7
6
5
(m = 0.912)
4

(32.69, 246.3)
2

100 9
8
7
6
5

10 2 3 4 5 6 7 8 9 2 3 4 5 6 7 8 9 2 3 4 5 6 7 8 9 2 3 4 5 6 7 8 9 2 3 4 5 6

0.1 1 10 100 1,000


Time (0 = 1175.566667)

Fig. 20—Log-log derivative and pressure difference for a bottomhole gauge with falling fluid level.

May 2015 SPE Production & Operations 93

ID: jaganm Time: 16:22 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 94 Total Pages: 15

G-Function Analysis Closure Events


G Time BHP DP
Bottomhole Pressure (MPa) C Closure 19.866 22.463 2.751
1st Derivative dp/dG
G* dp/dG

0.9

4.5
27

4.0
0.8
24

3.5
0.7
21
Pressure (MPa)

3.0
0.6
12 15 18

G*f(p,G)
f(p,G)
2.5 2.0
0.5 0.4
1

0.3

1.5 1.0
14
9

0.

0.2
6

0.5
0.1
3

0.0
0.0
(a) 9.00 18.00 27.00 36.00 45.00 54.00 63.00 72.00 81.00 90.00
G Time

3500 28000 500


Analysis 1 Inj. Volume 4.488 m3
ISIP 25151 kPaa
Ddalum 1668.790 m
Fracture Closure Fracture gradient 15.07 kPa/m
3150 27000 450
Gc 21.529
t c 625.98 minutes
pc 22 419 kPaa
2800 26000 400

2450 25000 350


Semilog Derivative G dp/dG (kPaa)

First Derivative dp/dG (kPaa)


2100 24000 300
Fracture Closure
Gc 21.529

p (kPaa)
t c 625.98 minutes
1750 pc 22 419 kPaa 23000 250

1400 22000 200

1050 21000 150

700 20000 100


Fracture Closure
Gc 21.529
t c 625.98 minutes
350 pc 22 419 kPaa 19000 50
Semilog Derivative
Pdata
First Derivative
0 18000 0
0 5 10 15 20 25 30 35 40 45 50
(b) G-function time

Fig. 21—G-function diagnostic derivative plots from different software-analysis systems.

Phase Segregation and Gas Entry. During the falloff period, ported by a partial vacuum between the fluid level in the well and
while the well is sitting undisturbed, gas entry from the formation the sealed wellhead. The rate of fluid-level fall is then controlled
can occur. For the ideal case of a sealed wellbore under isother- more by the rate of vaporization of the wellbore fluid than by the
mal conditions, the volume of the gas bubble remains constant as reservoir transmissibility. Fig. 19 shows a typical pressure
it rises. With no mass transfer from the gas to the wellbore fluid, response for a bottomhole gauge in a well with a falling fluid
the moles of gas in the bubble remain constant; therefore, the bub- level.
ble pressure remains constant as it rises. If a single gas bubble The data for the first few hours of falloff accurately represent
floats from the perforations to the surface under these conditions, the ISIP, initial leakoff rate, and closure of the fracture. The
the surface pressure will rise to the original bottomhole pressure remaining data, covering approximately 1 week of shut-in, show
and the pressure at the perforations will double. In reality, leakoff almost no change in the pressure on the gauge because the fluid
from the well is not identically zero—the increased pressure gen- head above the gauge is nearly constant. Analysis of this part of
erated by the rising bubble causes an increase in leakoff rate, the the pressure falloff will give no useful reservoir or fracture infor-
gas temperature decreases somewhat during transit, and some gas mation. Fig. 20 shows the log-log plot for this falloff, and pro-
may dissolve in the wellbore fluid. Therefore, a very small gas vides a useful clue for identifying the falling-fluid-level case. As
bubble entering at the perforations can cause a large pressure the fluid level begins to drop, the pressure derivative (magenta
upset, as demonstrated in Fig. 17. Fig. 18 shows the effects of gas line in Fig. 20) exhibits a large negative slope. Once the fluid
entry on an actual treatment. level in the well stabilizes and begins to drop at a rate controlled
by vaporization of the fluid at the surface, the derivative assumes
Loss of Hydrostatic Head. The loss of the hydrostatic head a positive, nearly straight-line slope of approximately one-half.
effectively terminates a DFIT. Bottomhole gauges can be and are During this period, the fluid at the surface of the standing column
installed to acquire data in certain circumstances, but these will is effectively boiling (fluid pressure equals vapor pressure), and
not aid in extending the test if the hydrostatic column starts to the heat of phase transformation causes the fluid temperature to
fall. In these cases, the bottomhole gauge will record the weight drop. This increases fluid density, driving a convection cell that
of the fluid standing above the gauge. The fluid column is sup- replaces the colder surface fluid with warmer fluid from below.

94 May 2015 SPE Production & Operations

ID: jaganm Time: 16:22 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 95 Total Pages: 15

Log-Log Analysis Closure Events


dT BHP DP
ΔP (MPa) C Closure 679.162 22.463 2.751
1st Derivative dDP/dDT BL Begin Linear Flow 4720.989 19.805 5.410
DT* dDP/dDT
EL End Linear Flow 12853.794 18.941 6.273

5 6 7
C BL EL

4
3
2
Pressure (MPa)
00
0.5

(a)
100.00 1000.00 10000.00
dT (minutes)

105
Inj. Volume 4.488 m3
6 ISIP 25 151 kPaa
Ddatum 1668.790 m
4 Fracture gradient 15.07 kPa/m
3
Δp, Semilog Derivative (Δt)dΔp/d(Δt)(kPaa)

104
6
4
3
2

103
6
4
3
2 Slope –1/2

2
10
6
4
3
2 Δpdata
Derivativedata
101
10–3 2 3 4 5 6 10–2 2 3 4 5 6 10–1 2 3 4 5 6 1.0 2 3 4 5 6 101 2 3 4 5 6 102 2 3 4 5 6 103 2 3 4 5 6 104
(b) Δ t (hours)

Fig. 22—Log-log pressure change vs. time after shut-in, with semilog derivative.

The important point here is that analysis of the pressure/time data to disparage any particular software or analyst, and any identify-
in this case cannot provide useful reservoir or fracture data. ing information about the specific test has been hidden or
There are three practical ways to avoid this problem. First, use removed whenever possible. The test data were acquired with a
a fluid of low enough density so that hydrostatic wellbore head is bottomhole gauge, and the test was pumped through the toe perfo-
less than reservoir pressure. This can be achieved in some cases rations in a horizontal well. Figs. 21a and 21b show the G-func-
through the use of oil (a surface gauge can then be used) if the res- tion diagnostic derivatives for the test, computed in two different
ervoir is not badly underpressured. Second, use a bottomhole software packages. The results are essentially the same, with clo-
gauge and bottomhole shut-in or isolation tool (that is fully opera- sure pressures of 22.463 and 22.419 MPa, respectively. The semi-
tional). The example in Figs. 19 and 20 was run with a bottomhole log derivative shows variable storage and a long after-closure
shut-in tool and memory gauge, but the shut-in tool leaked. The transient period. This part of the analysis is consistent across
third method is to vent the wellhead when the surface pressure almost all software. The slight difference in the value of G at clo-
drops to zero. This will allow the fluid level to fall to a pressure in sure may be because of slight differences in the computed pump-
equilibrium with the reservoir. Fracture closure will often occur ing time.
before the wellhead pressure drops to zero, and the closure time Once ISIP and closure are identified, the net extension pres-
and pressure will be within acceptable accuracy. After closure, sure can be computed from the difference. In this case, the ISIP is
there will be an extended transient period as the fluid level seeks slightly different because of wellbore-expansion effects of short
equilibrium. After that, the pressure decay to reservoir conditions duration. The next step in the analysis is the log-log plot of pres-
will be analyzable, but all reservoir transients will be delayed. sure change from ISIP vs. shut-in time. Figs. 22a and 22b show
Corrections can be made for the added injection volume in the af- these plots from the two analysis packages.
ter-closure calculations. After the end of the variable storage period, and for a very
brief time before closure, the plot in Fig. 22a shows a positive
Common Mistakes in After-Closure Analysis. The following one-half slope derivative corresponding to a linear fracture leakoff
example is one of many and shows some extremely common mis- mechanism. After closure, the derivative falls on a negative slope
takes in DFIT after-closure analysis. The example is not intended that is slightly shallower than negative one-half. Through the

May 2015 SPE Production & Operations 95

ID: jaganm Time: 16:22 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 96 Total Pages: 15

ACA-Linear Analysis Linear-Flow Events


BHP (MPa) FL Time (hours) BHP
BL Begin Linear 0.249 78.842 19.805
EL End Linear 0.149 214.389 18.941

EL BL

23.0 24.0
Bottomhole Pressure (MPa)
21.0 22.0
)
Pa
(M
84
7.55
19.0 20.0

re s1
ssu
Pre
re
Po
17.0 18.0

(a) 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00
Linear Flow Time Function

20200

19852

19504 Analysis 2
kh 0.53 md-m
k 0.1061 ma
p* 18026 kPaa
p (kPaa)

19156

18807

Δt 239.65 hours
P 18668 kPaa
18459

Pdata
18111
0.016 0.014 0.012 0.010 0.008 0.006 0.004 0.002
(b) 2
FR

Fig. 23—Pressure vs. linear and radial time functions for linear extrapolation to reservoir pore pressure.

entire test, the flow regime does not develop into an ideal pseudo- of the linear-flow time function. Entering the wrong pore pressure
linear-flow regime, and certainly shows no evidence whatsoever changes the shape and position of the pressure-difference curve,
of a pseudoradial-flow regime. The plot in Fig. 22b identifies the but does not alter the shape and position of the derivative curve.
late-time slope at negative one-half (shown in the green text box), Figs. 24a and 24b show the two after-closure flow-regime
which is an approximation. plots for the same test data. Fig. 24a shows the plot with the pore
The next step in the analysis should be to extrapolate the after- pressure estimated from the linear-flow extrapolation. The straight
closure pressure decline to estimate reservoir pore pressure. line connecting the beginning and end of the assumed linear-flow
Because the pseudolinear-flow regime is not fully developed, the period has the required slope of one-half. Clearly, the derivative
extrapolation of the Cartesian linear-flow plot will probably give falls on a significantly lower slope, which is consistent with the
a pressure estimate that is slightly too high. Use of the Cartesian slope of the semilog derivative in Fig. 23a. The pressure-differ-
radial-flow plot is not warranted at all. Fig. 23a shows the con- ence curve is forced to fall on the one-half slope and be separated
struction and extrapolation of the Cartesian linear-flow plot, with from the derivative by approximately two times because of the
an extrapolation to a pore pressure of 17.55 MPa. Fig. 23b shows selection of the pore-pressure value derived from the extrapola-
the construction and extrapolation of the Cartesian radial-flow tion of the Cartesian linear-flow plot (Fig. 23a).
plot that was presented in the other analysis. This extrapolation Fig. 24b shows the plot with the pore pressure picked from the
gives a reservoir pore pressure of 18.03 MPa. Cartesian radial-flow extrapolation. This value forces the pres-
In this case, the difference in the estimated pressure is prob- sure-difference curve to approach a –1 slope, and to fall on top of
ably insignificant (approximately 3%), although in many cases the the derivative curve. Clearly, the derivative is not on a –1 slope,
same error in flow-regime identification can lead to pressure and is still, in fact, of a slope less than negative one-half. There
errors of up to 30%. A more-serious error develops if the wrong appears to be a consistent trend in the industry to force a pseudor-
flow regime and wrong pore pressure are used in the Horner or adial solution out of a DFIT analysis, even when the data clearly
radial-flow solution for reservoir transmissibility. The best way to show that no such flow regime exists. The final results of this error
ensure that the after-closure flow regime is identified correctly is in flow-regime identification and the forcing of an imposed incor-
to use the Talley et al. (1999) log-log flow-regime plot. This plot rect analysis are most obvious in the computed reservoir transmis-
displays the difference between calculated bottomhole pressure sibility. Fig. 24b indicates an estimate permeability of more than
and static reservoir pressure (p* or pi) as a function of the square 0.1 md, which is consistent with a transmissibility of more than

96 May 2015 SPE Production & Operations

ID: jaganm Time: 16:22 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 97 Total Pages: 15

ACA Log-Log Analysis


Flow Events
P–Pi (MPa) FL2 Log Dir (P–Pi)
Log Derivative BL Begin Linear 0.061 0.957 2.246
EL End Linear 0.022 0.750 1.382

4 5 6
EL BL

3
2
Pressure (MPa)
1
0.5
pe =
w Slo
r Flo
Linea

0.1

0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10 0.20 0.30 0.40 0.50
(a) Squre Linear Flow (FL2)

104
7
5
4
3
Δp, Semilog Derivative (FL22 ) dΔp/d(FL22 ) (kPaa)

103
7
5
4
3 Impulse Radial -1
k 0.1061 md
2

102
7
5
4
3

2
Δpdata
Derivativedata

2 1.0 9 8 7 6 5 4 3 2 10–1 9 8 7 6 5 4 3 2 10–2


(b)
FL22

Fig. 24—Log-log of well pressure minus static reservoir pressure vs. the square of the linear-flow time function.

0.8 mdm/cp. These results can be reproduced in the software must also be met for the Horner extrapolation to be valid. In prac-
used for the Case (a) results when the same mistakes are made. tical terms, that means that extrapolation of an apparent straight
The closure time of G ¼ 20 from Fig. 21 is more consistent with a line at the end of a Horner plot will almost always be wrong.
system permeability of less than 0.002 md and transmissibility of Both the Cartesian radial-flow analysis and Horner plots can be
less than 0.014 mdm/cp. extrapolated to the pore pressure derived from a pseudolinear-
Applying the other relations described in this paper to this flow analysis (if the pseudolinear-flow regime exists), to obtain a
example gives a clearer interpretation of a consistent result. The maximum upper bound of reservoir transmissibility. Be aware
average pump rate for the test was 0.6 m3/m (3.7 bbl/min), with a that this estimate may be at least one order of magnitude too high.
pump time of 9.5 minutes. Fluid efficiency estimated from closure
time is 90%, with a transverse storage factor of 91%. On the basis
of Fig. 12, this should generate a fracture half-length of approxi- Conclusions
mately 100 m (320 ft). The rule of thumb for closure from Fig. 13, Diagnostic fracture injection tests (DFITs) can provide a wide va-
assuming a permeability of approximately 0.002 md, is 10 hours. riety of data to aid in designing hydraulic-fracturing treatments
Actual closure time was 688 minutes (11.4 hours). A valid linear- and characterizing the subject reservoir. However, there are many
flow reservoir transient may be expected after 30 hours, but a fully factors associated with data collection and analysis that can result
developed pseudolinear-flow regime never developed, even up to in poor or incorrect results. This paper attempts to describe some
the end of the test at 330 hours (13.8 days). According to Fig. 15, a of the common problems and to help prevent some common
pseudoradial-flow regime consistent with a single planar fracture errors often observed in DFIT execution and analysis. These
should not be expected before 300 days. Clearly, any results issues (and their possible solutions) include
obtained from a forced radial-flow analysis are not valid. • DFITs should be pumped with a Newtonian, nonwall-building
A final comment about the use of the Horner plot in after- fluid.
closure analysis is warranted. The Horner plot is only valid, and • Multiple pump-ins must be avoided.
should only be relied upon, when a fully developed pseudoradial- • The tests must be planned to allow for closure and enough fall-
flow regime is identified in the data. The conditions described in off time to enter linear flow at a minimum. This time period
the preceding for the use of the Cartesian radial-flow analysis may be several days, if not weeks, in low-permeability systems.

May 2015 SPE Production & Operations 97

ID: jaganm Time: 16:22 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030


PO169539 DOI: 10.2118/169539-PA Date: 8-May-15 Stage: Page: 98 Total Pages: 15

• Use stepdown, not stepup, tests to determine near-wellbore and Cinco-Ley, H. and Samaniego-V., F. 1981. Transient Pressure Analysis
perforation friction. for Fractured Wells. J Pet Technol 33 (9): 1749–1766. SPE-7490-PA.
• Correct instantaneous-shut-in-pressure determination is critical http://dx.doi.org/10.2118/7490-PA.
to further analysis of the test. Wellbore compressibility and Craig, D. P., Eberhard, M. J., Ramurthy, M., et al. 2005. Permeability,
other factors must be taken into consideration. Pore Pressure, and Leakoff-Type Distributions in Rocky Mountain
• The tests must be of adequate duration and rate (shortest practi- Basins. SPE Prod & Oper 20 (1): 48–59. SPE-75717-PA. http://
cal duration at highest practical rate). dx.doi.org/10.2118/75717-PA.
• The possibility of phase segregation and gas entry must be Craig, D.P. and Blasingame, T.A. 2006. Application of a New Fracture-
considered. Injection/Falloff Model Accounting for Propagating, Dilated, and
• Loss of hydrostatic head will likely render the test useless, even Closing Hydraulic Fractures. Presented at the SPE Gas Techology
with bottomhole gauges, unless certain additional steps are Symposium, Calgary, 15–17 May. SPE-100578-MS. http://dx.doi.org/
taken. 10.2118/100578-MS.
• Not all tests result in pseudoradial-flow regimes, and, in fact, Craig, D.P. and Brown, T.D. 1999. Estimating Pore Pressure and Perme-
reaching pseudoradial conditions is more the exception than the ability in Massively Stacked Lenticular Reservoirs Using Diagnostic
rule. There are numerous reasons for false indications of pseu- Fracture-Injection Tests. Presented at the SPE Annual Technical Con-
doradial conditions. To avoid these, full integration of the test ference and Exhibition, Houston, 3–6 October. SPE-56600-MS. http://
results must be considered and the physics of the solution must dx.doi.org/10.2118/56600-MS.
be respected. Daltaban, T. S. and Wall, C. G. 1998. Fundamental and Applied Pressure
Consideration of these points and the others listed in this paper Analysis, Chap. 15.8, 438. London, England: Imperial College Press.
will lead to improved data acquisition and analysis, which in turn, Nolte, K. G. 1979. Determination of Fracture Parameters from Fracturing
will lead to better tests, better results, and ultimately better predic- Pressure Decline. Presented at the SPE Annual Technical Conference
tions of production and reserve recovery. and Exhibition, Las Vegas, Nevada, 23–26 September. SPE-8341-MS.
http://dx.doi.org/10.2118/8341-MS.
Talley, G.R., Swindell, T.M., Waters, G.A. et al. 1999. Field Application
Nomenclature
of After-Closure Analysis of Fracture Calibration Tests. Presented
d ¼ pipe diameter, L, in. at the SPE Mid-Continent Operations Symposium, Oklahoma City,
E ¼ Young’s modulus, psi, m/Lt2 Oklahoma, 28–31 March. SPE-52220-MS. http://dx.doi.org/10.2118/
Esteel ¼ Young’s modulus of steel, psi, m/Lt2 52220-MS.
H ¼ fracture height, ft, L Tompkins, D., Waldman, N., Redman, M. et al. 2014. Technology Update
k ¼ permeability of the leakoff system, md, L2 1: Effects of Ambient Temperature Change on Diagnostic Fracture
k1,2,3 ¼ permability in series, md, L2 Injection Testing. JPT 66 (9). http://www.spe.org/jpt/article/7119-
kavg ¼ average permability of a series, md, L2 technology-update-1-4/.
L ¼ fracture length, ft, L Willhite, G.P. 1986. Waterflooding. SPE Textbook Series Vol. 3, first edi-
L1,2,3 ¼ length, ft, L tion. Richardson, Texas: Society of Petroleum Engineers.
LT ¼ total length of a series, ft, L Young, W. and Budynas, R. 2011. Roark’s Formulas for Stress and Strain,
Lthickness ¼ wall thickness, L, in. seventh edition. New York City, New York: McGraw-Hill Companies.
N ¼ moles of gas, lbm mol, m
Pinternal ¼ internal pressure change, psi, m/Lt2 Robert Barree is president and principal investigator of Barree
Phead ¼ hydrostatic head, psi, m/Lt2 & Associates. Previously, he was a senior technical consultant
DPT ¼ total pressure drop, psi, m/Lt2 at Marathon’s Petroleum Technology Center. Barree’s re-
P1 ¼ pressure at the perforations, psi, m/Lt2 search interests include well completions, stimulation, numeri-
P2 ¼ pressure at the surface, psi, m/Lt2 cal simulation, special core analysis, and formation damage.
qi ¼ injection rate in bbl/min, V/t He has authored or coauthored more than 70 technical
qi,max ¼ maximum injection rate in bbl/min, V/t papers. Barree holds a BS degree from Pennsylvania State Uni-
R ¼ real-gas constant, ft3psi/lbm mol- R versity and a PhD degree in petroleum engineering from the
t ¼ time, minutes, t Colorado School of Mines. He is a member of SPE.
tc ¼ time to fracture closure, minutes, t Jennifer Miskimins is a senior consulting engineer at Barree &
tp ¼ pumping time, minutes, t Associates. Previously, she was an associate professor at the
T ¼ temperature,  R, T Colorado School of Mines and held various positions at Mara-
W ¼ fracture width, in. thon Oil Company. Miskimin’s research interests include hy-
z ¼ compressibility factor draulic fracturing, completions, and unconventional-reservoir
development. She has authored or coauthored more than 65
technical papers. Miskimins holds a BS degree in petroleum
engineering from Montana Tech and MS and PhD degrees in
References
petroleum engineering from the Colorado School of Mines.
Barree, R.D. 1998. Applications of Pre-Frac Injection/Falloff Tests in Fis- She is a member of SPE.
sured Reservoirs – Field Examples. Presented at the SPE Rocky
Mountain Regional/Low-Permeability Reservoirs Symposium, Den- John Gilbert is a senior engineering consultant at Barree &
Associates. Previously, he worked for Santos in Australia as a
ver, 5–8 April. SPE-39932-MS. http://dx.doi.org/10.2118/39932-MS. stimulation team leader and for Marathon Oil Company in the
Barree, R.D., Barree, V.L, and Craig, D.P. 2009. Holistic Fracture Diag- Drilling and Completions group. Gilbert’s research interests
nostics: Consistent Interpretation of Prefrac Injection Tests Using Mul- include production analysis, well testing, and fracture-stimula-
tiple Analysis Methods. SPE Prod & Oper 24 (3): 396–406. SPE- tion modeling. He holds an ME degree in petroleum engineer-
107877-PA. http://dx.doi.org/10.2118/107877-PA. ing from Colorado School of Mines. Gilbert is a member of SPE.

98 May 2015 SPE Production & Operations

ID: jaganm Time: 16:22 I Path: S:/3B2/PO##/Vol00000/140030/APPFile/SA-PO##140030

Вам также может понравиться