You are on page 1of 9

SPE 134534

Practical use of well test deconvolution

Gringarten, A. C., SPE, Imperial College London

Copyright 2010, Society of Petroleum Engineers

This paper was prepared for presentation at the 2010 SPE Annual Technical Conference and Exhibition held in Florence, Italy, 20–22 September 2010.

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
Deconvolution transforms variable rate pressure data into a constant rate initial drawdown with a duration equal to the total
duration of the test. It yields directly the corresponding pressure derivative, normalized to a unit rate. It is not a new
interpretation method, but a new tool to process pressure and rate data in order to obtain more pressure data to interpret with
conventional techniques.

Although deconvolution has received much attention since 2001 following the publication of a stable algorithm by von
Schroeter, Hollaender, and Gringarten, its use by practicing engineers is still limited. One reason is limited access to the
algorithm, which had only recently become available in commercial well test analysis software products. The other is concerns
on how deconvolution should be used and how reliable it is. As a result, few examples of practical applications of well test
deconvolution are available in the literature.

This paper illustrates various uses of deconvolution in tests of short and long durations. Examples include DST’s with
erroneous rates, which deconvolution is able to correct; and data from permanent downhole pressure gauges in vertical and
horizontal wells, where deconvolution shows compartmentalization and recharge from other layers, which could not be seen in
the original data.

Recommendations on how to perform deconvolution and how to verify deconvolution results are also provided. It is hoped
that this paper will encourage well test interpreters to use deconvolution confidently as part of the well test analysis process.

Introduction
Deconvolution has received much attention in the past decade (von Schroeter et al. 2001, 2002, 2004; Gringarten et al.
2003; Levitan 2003; Levitan et al. 2004; Ilk et al. 2005, 2006; Amudo et al. 2006; Gringarten 2006; Whittle et al. 2008, 2009;
Aluko and Gringarten 2009), following the publication of a stable deconvolution algorithm (von Schroeter et al. 2001). It is
not a new interpretation method, but a new tool that processes pressure and rate data to obtain more pressure data to interpret.
Deconvolution transforms variable-rate pressure data into a constant-rate initial drawdown with a duration equal to the total
duration of the test, and yields directly the corresponding pressure derivative, normalized to a unit rate. This derivative is
therefore free from distortions caused by pressure-derivative calculation algorithms, and from errors introduced by incomplete
or truncated rate histories (Gringarten 2006). It also exhibits all the flow regimes that have been dominating throughout the
test.

The Schroeter et al. (2001) deconvolution algorithm, based on the Total Least Square method, has been implemented, with
some variations, in most commercial well test analysis softwares. Deconvolution is therefore readily available to all well test
interpreters. Yet, by and large, its use seems to be mostly restricted to experts. One obvious reason is that the method is still
new, and all new techniques take some time before they are adopted by practicing engineers (the well known “chasm”, Moore
1991). This is not different from what happened when pressure type curve analysis was first introduced (Ramey 1970;
McKinley 1971): it was suggested that it should be used only in emergency, after more conventional techniques had failed
(Earlougher, 1977). Another reason is that the theoretical limitations of deconvolution and the difficulties that could be
encountered using the deconvolution process have been over-emphasized. Moreover, shortcomings of particular software
implementations have often been presented as limitations of deconvolution itself (as with derivatives in the mid-1980’s).

This paper illustrates various uses of deconvolution in tests of short and long durations. Examples include DST’s with
erroneous rates, which deconvolution is able to correct; and data from permanent downhole pressure gauges in vertical and
horizontal wells, where deconvolution shows compartmentalization and recharge from other layers, which could not be seen in
2 SPE 134534

from conventional analyses. Recommendations on how to perform deconvolution and how to verify deconvolution results are
also provided. It is hoped that this paper will encourage well test interpreters to use deconvolution with confidence as part of
their well test analysis process.

The deconvolution process


Algorithm
The deconvolution algorithm estimates both rates (called “adapted” rates herein) and rate-normalised derivative by
minimising an error measure, E, which is a weighted combination of a pressure match, a rate match, and a penalty term based
on the overall curvature of the graphed derivative, whose purpose is to enforce derivative smoothness (Eq. 1):
E= ( pi − y ∗ g ) − p 2
2
+ ν ( y − q ) 2 + λ (Dz − k )
2 2
2
(1)
In Eq. 1, p and q are the measured pressure and rate data, respectively; pi is the initial pressure at the start of the rate history,
which can be an input into or an output from deconvolution; D z − k is the curvature measure; λ and ν are weights for the
2
2
rate match and the curvature, normalized to the weight for the pressure match (different well test analysis softwares define λ
and ν differently); y is the adapted rate; and g is the instantaneous source function (Gringarten and Ramey, 1973), i.e. the
derivative of the pressure with respect to time. Both y and g are outputs of deconvolution.
Actually, t.g(t), the derivative of the pressure with respect to the natural log of time, is calculated instead of g(t): an
arbitrary shape of t.g(t) is input as an initial guess into Eq.1, and it is modified in successive iterations until the error mesure E
is minimized (Fig. 1). The result is the deconvolved derivative.
Deconvolution is in principle only valid for linear systems, i.e. reservoirs with a slightly compressible fluid (in practice,
deconvolution can be used with pseudo-pressures in the case of gas or multiphase flow). A further assumption is that the
pressure is uniform in the reservoir at the start of the rate history.
Deconvolution could be performed on a single flow period1, several flow periods, or part or all of the pressure history. The
resulting deconvolved derivative is defined piecewise: in the case of a single flow period, it reproduces the flow period
derivative at early times (as it would appear for the very first drawdown in the reservoir after pressure stabilisation). At later
times, it is defined only from the start to the end of the flow period, as measured from the start of the test. It is interpolated
between the two. In the case of multiple flow periods, on the other hand, the early part of the deconvolved derivative is some
average of the derivatives for each flow period, and the later part is defined only from the start to the end of each flow period.
The complete deconvolved derivative is again obtained by interpolation. The “average” at early times may be different from
any of the original flow period derivatives, if changes have occured, for instance, in skin effects or wellbore storage
coefficients. It could also be polluted by noise from some of the deconvolved flow periods, such as phase redistribution in the
wellbore. This is not too much of a concern, as the primary objective of deconvolution is to generate data beyond what is
available in individual flow periods (yellow bands in Fig. 2). Data already available in a particular flow period can be analysed
by conventional means.
In practice, one tends to deconvolve individual build ups, or groups of build ups, or continuous ranges of pressure history.
Drawdowns are often too noisy to be deconvolved individually.

Pressure history Pressure history


4000 4000
FP19

3500 3500
Pressure, psia
Pressure, psia

FP203 FP203
3000 3000

2500 2500

2000 2000

1500 1500

1000 1000
log [t g(t)]

10-3 10-1 10 103 105 10-3 10-1 10 103 105


Time from the start of the test, hrs Time from the start of the test, hrs
Iteration 1
3
FP Rate FP Pressure λ Initial pressure FP Rate FP Pressure λ Initial pressure
5
7 #(1-203)[203]{6.64424E+05}3696.00 #(1-203)[19, 203]{1.21662E+05}3696.00
102 102
Deconvolved derivative
mn(p) Derivative, psi
mn(p) Derivative, psi

10 10

1 1

10-1 FP203 10-1 FP203


FP19

10-2 10-2
log ∆t 10-3 10-1 10 103 105 10-3 10-1 10 103 105
Elapsed time, hrs Elapsed time, hrs

Figure 1: Iterative calculation of deconvolved derivative Figure 2: Well 1: Construction of the deconvolved derivative

1
In this paper, a flow period (FP) refers to a period at constant rate, which could be a drawdown or a build up.
[SPE 134534] 3

The construction of the deconvolved derivative is illustrated in Fig. 2 with test data from Well 1 (Figs. 3 and 4). Well 1
has been drilled horizontally in the upper layer (Zone 1) of a two-layer gas condensate reservoir with a two-meter thick shale
layer separating Zone 1 from Zone 3 (Fig. 3). Deconvolved derivatives in Fig. 2 and in this paper are identified by a label
which describes the conditions of the deconvolution:
• (…..) refers to the rate record used. For instance, (1-203) means that the rate history with 203 different rates has been
used.
• […..] identifies the pressure data that have been deconvolved: [19] and [203] means that only the data from FP19 and
FP203 were involved. [1-203] corresponds to the deconvolution of the entire pressure history up to the end of FP203
(Figs. 6 to 9).
• {….} states the value of the regularisation parameter λ
• The last parameter in the label represents the initial pressure (pav)i
The left hand side of Fig. 2 shows the deconvolved derivative for a single flow period, FP203, a 1785 hour build up from
14500 hours to 16384 hours of production. The vertical axis is labeled in terms of normalized pseudo-pressure, a modification
of the single-phase pseudo-pressure function used to linearize the diffusivity equation in gas reservoirs (Meunier et al. 1987).
The deconvolved derivative is defined by the derivative of FP203 from 0 to about 100 hours, and by FP203 between 14500
hours and 16384 hours. The difference between the deconvolved derivative and the actual FP203 derivative in the 100 to 1785
hour range is due to the fact that the former is a drawdown derivative, corresponding to the first drawdown after stabilization,
whereas the latter is a multirate derivative, which is affected by the rate history when boundaries have been reached.
The right hand side of Fig. 2 shows the deconvolved derivative for both FP19 and FP203. FP19 is an 11 hour build up between
809 hours and 820 hours, during the DST. The deconvolved derivative is a combination of the FP19 and FP203 derivatives
from 0 to 1785 hours, and is defined by FP19 between 809 hours and 820 hours and by FP203 between 14500 hours and
16384 hours. In this particular example, FP19 and FP203 do not overlap: FP19 is infinite acting and dominates at early and
middle times, whereas FP203 has boundary effects and dominates at late times. The yellow bands in Fig. 2 represent the
increase in the radius of investigation, one full log-cycle in this case.

Well 1 schematic

Well (L=900m) Well 1: Production


Zone 2 100m 2m 4000 200
Shale layer FP 35
FP 66
Zone 3 70m Pressure 180
FP 120 FP 226
Dew point pressure Measured rates
FP 186
Analysis rates 160
FP 203
3000 FP 235
Well 1 : DST 140
3700 100

Total Rate (MMscf/D)


Pressure (psia)

FP 198 120
FP 5 FP 15 FP 19 FP 218
3660 80 FP 216
2000 FP 386 100
Total rate (MMscf/D)

FP 428
Pressure (psia)

FP 465 80
3620 FP 10 60

FP 389 FP 440 60
3580 40 1000 FP 480 (Dd)
40

3540 Dew point pressure 20


20

3500 0 0 0
730 750 770 790 810 830 0 10000 20000 30000 40000 50000 60000
Elapsed time (hrs) Elapsed time (hrs)

Figure 3: Well 1: DST Figure 4: Well 1: production history

Control parameters
Deconvolution weight parameters
The estimate E in Eq. 1 depends on two weights, ν for the rate match, and λ for the roughness penalty. ν is set at a default
N ∆p 2
2

value, vdef = , where ∆p is the vector of pressure drop data, q is the vector of measured rates, m is the number of
m q2
2
pressure points and N is the number of flow periods (von Schroeter et al. 2004). In the case of the deconvolution of a single
∆p
2

flow period, λ is also set at a default value, λdef = 2 . For multiple flow period deconvolution, λ must usually be set at 10
m
to 100 times λdef to avoid oscillations on the derivative. Regularisation introduces bias, and thus the user must choose a level
of λ that imposes just enough smoothness to eliminate small-scale oscillations on the deconvolved derivative while preserving
genuine reservoir features. This is illustrated in Fig. 5 for the deconvolution of Well 1 pressure history from the start of the test
4 SPE 134534

to the end of FP203. Oscillations that exist at middle times in the deconvolved derivative for the λdef value of 4.41231.104
increase as λ decreases to λdef /10 and decrease as λ increases to 10 λdef and 100 λdef. All the derivatives converge at late times,
suggesting a satisfactory deconvolution in all cases.

100

#(1-203)[19,203]{1.21662E+05}3696.00 1-19
#(1-480)[19]{6.22016E+01}3696.00 #(1-480)[19]{8.32792E+00}3690.00
1-19
#(1-203)[1-203]{4.41231E+06}3696.00 BEST 1-19
#(1-480)[19]{4.83274E+01}3694.00 #(1-480)[19]{3.64133E+01}3693.00
1-19
10 #(1-203)[1-203]{4.41231E+05}3696.00 1-19
#(1-480)[19]{2.63689E+01}3692.00 #(1-480)[19]{1.83950E+01}3691.00
1-19
#(1-203)[1-203]{4.41231E+04}3696.00 DEFAULT 1-19
#(1-480)[19]{1.23814E+01}3697.00 #(1-480)[19]{7.80361E+01}3698.00
10 1-19
#(1-203)[1-203]{4.41231E+03}3696.00
1-19
#(1-480)[19]{1.15795E+02}3700.00
Deconvolved Derivative

Deconvolved Derivative
1

#(1-203)[203]{6.60713E+05}3690.00
1 #(1-203)[203]{6.54761E+05}3696.00
0.1
#(1-203)[203]{6.64424E+05}3700.00

0.01

0.1

0.001

0.01
0.0001
0.001 0.01 0.1 1 10 100 1000 10000 100000
0.001 0.01 0.1 1 10 100 1000 10000 100000
Elapsed Time hrs
Elapsed Time hrs

Figure 5: Impact of λ on deconvolution of an entire pressure Figure 7: Impact of pi on deconvolution


history

Initial pressure
The deconvolved derivative is very sensitive to the initial pressure if the flow period being deconvolved is infinite acting, and
not sensitive if boundaries have been reached. This is illustrated in Fig. 7, which shows different shapes of FP19 deconvolved
derivatives for different values of pi. The FP203 derivatives, on the other end, are not affected. A good estimate of the initial
reservoir pressure is therefore required, which can be obtained from a preliminary analysis of the data. Alternatively, pi can be
obtained by trial and error if at least two different, infinite acting build ups are available, as a correct pi must yield the same
deconvolved derivative (Levitan et al., 2004). This is shown in Fig. 8 for FP15 and FP19, the last two DST build ups: an initial
pressure of 3696 psia yields almost identical deconvolved derivatives.

102

Pressure Difference (psi), Measured - simulated


100
FP Rate FP Pressure λ Initial pressure 80
60

FP15 #(1-15)[15]{3.69371E+01}3696.00 40
20
0
10 FP19 #(1-19)[19]{3.31900E+01}3696.00
-20
mn(p) Derivative, psi

-40
-60
-80
-100
3800
1
Data #(1-203)[1-203]{4.41231E+06}3696.00 BEST
3600 #(1-203)[1-203]{4.41231E+05}3696.00
#(1-203)[1-203]{4.41231E+04}3696.00 DEFAULT
Pressure (psia)

3400 #(1-203)[1-203]{4.41231E+03}3696.00

10-1 3200

3000
FP202
2800
10-2
10-3 10-2 10-1 1 10 102 103 104 105 2600
0 2000 4000 6000 8000 10000 12000 14000 16000 18000
Elapsed time, hrs Elapsed time (hrs)

Figure 8: Well 1: Verification of pi Figure 9: Well 1: Pressure history match

Verification of deconvolution
Deconvolution quality is assessed by the lack of late time oscillations on the deconvolved derivatives; by the fact that
different flow periods yield converging derivatives (as in Fig. 5); and by the consistency between the deconvolved derivatives
and possible interpretation models resulting from a conventional analysis of the data. A further and most important check is
the ability to generate a pressure history from the deconvolved derivative that matches the actual pressure history. Such a
match is shown in Fig. 9 for the deconvolved derivatives of Fig. 5. All deconvolved derivatives yield a satisfactory match with
actual pressure data, with maximum errors of less that 10% in the drawdowns.

Interpretation methodology using deconvolution


The methodology for well test analysis using deconvolution is illustrated in Fig. 10. Pressure and rate data are deconvolved
using appropriate values of the weight parameters ν and λ. For a gas well, pressure data are converted to pseudo-pressure or
normalised pseudo-pressure in order to approximate a linear system before applying deconvolution. Once a satisfactory
derivative has been obtained, a convolved pressure history is calculated from that derivative, the adapted rates and the initial
pressure obtained from or used in deconvolution, and compared with the measured pressure history. If the match is acceptable,
[SPE 134534] 5

the deconvolved derivative is used to generate a unit-rate pseudo-pressure drawdown which has the same duration of that of
the entire test. This unit-rate drawdown is analysed in the conventional way. The resulting model is then applied to the
measured pressure data using the adapted rates, and the model parameters are refined until an acceptable match is obtained.
0.1

Unit rate Deconvolved Normalised Pseudo-Pressure Derivative


Pressure data p Rates FP Rate FP Pressure λ Initial pressure
Slope > 1
mn(p) vs. p #(1-66)[19, 66]{2.03480E+04}3696.00
#(1-186)[19,186]{8.47784E+04}3696.00
Pseudo-pressure mn(p) #(1-203)[19,203]{1.21662E+05}3696.00
0.01
Deconvolve #(1-386)[19,386]{3.80886E+05}3696.00
#(1-465)[19,465]{1.37137E+05}3696.00
#(1-480)[19-480]{1.00000E+06}3696.00
Deconvolved mn(p) derivative Adapted rates

Single layer closed reservoir behaviour


Convolve
0.001

Convolved pseudo-pressure with adapted rates

NO
Compare

YES
0.0001
Unit rate Convolve

Unit rate convolved pseudo-pressure drawdown Adapted rates

Unit rate convolved


Multilayer layer closed reservoir behaviour
p vs. mn(p) Interpretation Refine
pressure drawdown Analyse Convolve END
model match
0.00001

Pressure data 0.001 0.01 0.1 1 10 100 1000 10000 100000

Elapsed Time hrs

Figure 10: Deconvolution interpretation methodology (Amudo Figure 11: Well 1 deconvolved derivatives
et al. 2006)

Identification of boundaries and connectivities not visible by conventional analyses


One primary advantage of deconvolution is in the increase of the radius of investigation, which allows seeing reservoir
features not visible in individual flow periods. Examples can be found in Gringarten (2006) and Aluko and Gringarten (2009).
This is further illustrated here with Well 1 in Figs. 11 to 14, using the methodology described by Fig. 10. In the log-log plot of
Fig. 11, production build ups (FP66, 185, 203, 386, 465 and 480) are deconvolved together with FP19 to add early time data to
the deconvolved derivatives. All deconvolved derivatives exhibit unit slope log-log straight lines at late times, indicating a
closed reservoir2. The late time unit-slope log-log straight lines for the later build ups (FP386 and 465) merge into a single
straight line which is shifted to the right compared to the merged unit-slope log-log straight line for the first one year and a half
of production (FP66, 186 and 203), indicating an increased drainage volume.
Fig. 11 also shows the deconvolved derivative of all pressure data from FP19 to FP481 (the quality of this derivative is
verified in Fig. 12). At late times, it first follows the unit-slope straight line corresponding to FP66, 186 and 203, and then
trends toward that for FP386 and 440. This suggests a multilayered behaviour due to drainage of Zone 3 through the shale
layer. The transition occurs around 5000 hours and cannot be seen on any of the individual build ups or drawdowns.
The unit-rate drawdown convolved from the FP19-480 deconvolved derivative is analysed with a 3-layer closed reservoir
model in Fig. 13. The top, middle and bottom layers represent Zone 2, the shale layer, and Zone 3, respectively. The FP19 to
481 multilayer behaviour is matched with a shale layer vertical permeability of 10-4 mD. The reservoir is a closed rectangle,
with boundaries at distances of approximately 300m, 2200m, 400m and 800m.
As a final step, the multilayer model is applied to actual production build ups and the corresponding analysis parameters are
adjusted as required to improve the match (Fig. 14).
3700 2200

Gas Rate (Mscf/D)


Pressure (psia)

3680 1800
3800
3660 1400
(1-480)[19-480]{1.00000E+06}3696
3600
3640 1000
Actual pressure
3400 600
3620
200
3200 3600
0 10000 20000 30000 40000 50000 60000
Elapsed time (hrs)
3000
Model
nm(p) Change and Derivative (psi)

Log-Log Match Horner Match - Flow Period 2


103 3700 Uniform Flux Horizontal Well with C and S
Pressure (psia)

2800 Multi-layer (with cross-flow)


Rectangle
102 3690
Pressure (psia)

2600 Results
10 3680 (pav)i 3696 psia
(pav)f 3655 psia
1 3670 pwf 3655.4 psia
2400 (kh)t 15200 mD.ft
k (av) 23. mD
10-1 3660
L 900 m
2200 S(t) -5.2
10-2 3650 S(w) 3.1
10-3 10-1 10 103 105 107 109 -4000 -2000 0 2000 4000 S(c) -5.8
Elapsed time (hrs) Zw 51 m
2000 Superposition Function (Mscf/D) C 11 bbl/psi
k1 (xy) 26 mD
k2 (xy) 0.0001 mD
Simulation (Constant Skin) k3 (xy) 20 mD
1800 3700 k1 (z) 5 mD
k2 (z) 0.0001 mD
3690 k3 (z) 2.000 mD
S(1) -5.1
Pressure (psia)

1600 S(2) Non Perf.


3680 S(3) Non Perf.
d1(1:3) 290 m
d2(1:3) 2170 m
1400 3670
d3(1:3) 400 m
d4(1:3) 810 m
3660 Type d1(1:3) No Flow
1200 Type d2(1:3) No Flow
Type d3(1:3) No Flow
0 10000 20000 30000 40000 50000 60000 3650
Type d4(1:3) No Flow
0 10000 20000 30000 40000 50000 60000
Dp(S) 0.97 psi
Elapsed time (hrs)
Elapsed time (hrs)
Figure 13: Well 1: Analysis of the unit rate drawdown
Figure 12: Well 1: verification of FP19-480 deconvolution generated from FP19-480 deconvolved derivative

2
The straight line slope is not exactly one because of non-linearity due to gas flow, as observed by Levitan and Wilson (2010)
who propose a correction based on material balance. The slope is close enough to one, however, so that a closed system can be
diagnosed without ambiguity.
[SPE 134534] 6

Log-Log Match - Flow Period 186 Horner Match - Flow Period 186 Simulation (Constant Skin) - Flow Period 186
nm(p) Change and Derivative (psi)

1000 3200 4000

3100

Pressure (psia)
Pressure (psia)
100 3000

3000

10 2000

2900

1 1000
2800

0.1 2700 0
0.001 0.01 0.1 1 10 100 1000 10000 100000 0 100 200 300 400 500 0 10000 20000 30000 40000 50000 60000

Elapsed time (hrs) Superposition Function (MMscf/D) Elapsed time (hrs)

Figure 14: Well 1: analysis of build up FP186 with the interpretation model from the unit-rate drawdown analysis

Correction of rates
Another use of deconvolution is in correcting erroneous rates or determining missing rates. This is achieved by
deconvolving the entire rate history and adapting the entire rate history.

Test with erroneous rates


Rate correction is illustrated with the test for oil Well 2 shown in Fig. 15. The test includes a number of drawdowns and build
ups, with the two main build ups identified as FP12 and FP39. Rates are clearly not consistent with the pressure trend,
especially during the main drawdown between 320 and 375 hours. The rate inconsistencies are confirmed in the rate-
normalised log-log plot in Fig. 16. On this plot, the derivatives for the first and second build ups, FP12 and FP39, show
stabilisations after 5 hours which could correspond to radial flow. The different stabilisation levels could imply a variation of
permeability during the test. More likely, they indicate errors in the rate measurements. These rate errors can be corrected by
deconvolution.
5000 3 10000
Rate Normalised nm(p) Change and Derivative (psi)

4000
Pressure (psia)

Time-corrected Total Rate (MMscf/D)

3000

2000
2 1000
1000 FP 12 FP 39 FP 12
0
FP 39
1 100
Radial flow stabilizations?

0 10
0 100 200 300 400 500 0.01 0.1 1 10 100 1000
Elapsed time (hrs) Elapsed time (hrs)

Figure 15: Well 2: Pressure and rate histories Figure 16: Well 2: Rate normalized log-log plot

1000 1000
Deconvolved Derivative

Deconvolved Derivative

FP 12 FP 12
100 100
10000
FP 39 FP 39
#(1-12) [12]{1.47234E+06}3770 #(1-39) [39]{1.43754E+06}3770
10 10
#(1-39)[1-39]{2.38812E+08}3770 #(1-39)[1-39]{8.00000E+07}3770
BEST
#(1-39)[1-39]{1.00000E+08}3770 #(1-39)[1-39]{1.50000E+08}3770
1 1
Deconvolved Derivative, psia

#(1-12)[12]{1.49323E+06}3770 #(1-39)[39]{1.43754E+06}3770 #(1-12)[12]{1.72790E+06}3700 #(1-39)[39]{1.33163E+06}3700 1000


0.1
0.01 0.1 1 10 100 1000
0.1
0.01 0.1 1 10 100 1000
FP 1-39
Elapsed Time hrs Elapsed Time hrs

1000
FP 39
Deconvolved Derivative

FP 12
100 100
FP 39
FP 12
10

#(1-39)[12]{1.01449E+06}3520 #(1-39)[39]{9.74062E+05}3520 10
0.1 0.01 0.1 1 10 100 1000
0.01 0.1 1 10 100 1000
Elapsed time, hrs
Elapsed Time hrs

Figure 17: Well 2: Determination of pi Figure 18: Well 2; Deconvolution with pi = 3770 psia

The initial pressure for deconvolution is determined by trial and error, according to the procedure described by Levitan et al.
(2004). The deconvolved derivatives for the two build ups, calculated for a range of pi values based on RFT measurements, are
compared in Fig. 17: the deconvolved derivatives must coincide at late times for pi to be correct; otherwise, they diverge or
[SPE 134534] 7

cross. The best estimate from Fig. 17 (3770 psia) is then used to deconvolve the entire pressure history (FP1-39 in Fig. 18).
Several λ values are tried to obtain the “best” deconvolved derivative, which corresponds to λ = 1.5 108. It has a duration of
366 hours, which is the duration of the entire test, whereas the duration of the longest build up is 24 hours only. The quality of
the deconvolution is verified in Fig. 19 by comparing the pressure history calculated from the deconvolved derivative of FP1-
39 with the actual pressure measurements. The match is reasonably good for all the values of λ used in the deconvolution.

The rates corrected by deconvolution of FP1-39 are shown in Fig. 20 (continuous lines). Although they slightly differ for
different values of λ, they all show the same overall trend and are different from the reported rates (dashed line), except where
the rates have actually been measured. The corresponding rate-normalised log-log plot is shown in Fig. 16. Contrary to rate-
normalised log-log plot in Fig. 11, this plot now shows good consistency between FP12 and FP39.

4000
FP 12 FP 39
#(1-39)[1-39]{2.38812E+08}3770 #(1-39)[1-39]{8.00000E+07}3770
5000 #(1-39)[1-39]{1.00000E+08}3770 #(1-39)[1-39]{1.50000E+08}3770 3
3000
Pressure (psia)

Given rates
4000

Pressure (psia)
2000 #(1-12) [12]{1.47234E+06}3770
#(1-39)[1-39]{2.38812E+08}3770 3000
#(1-39)[1-39]{1.00000E+08}3770
1000 #(1-39) [39]{1.43754E+06}3770
2000
FP 12 FP 39

Total Rate (MMscf/D)


#(1-39)[1-39]{8.00000E+07}3770 2
#(1-39)[1-39]{1.50000E+08}3770
1000
0
0 100 200 300 400 500
Elapsed time (hrs)
0

3600 4000

3200 FP 12 FP 39 1
3000
2800
Pressure (psia)

Pressure (psia)

2400
2000
2000

1600
1000
1
1200 0
0 100 200 300 400 500
800 0 Elapsed time (hrs)
94 96 98 100 102 104 106 108 410 415 420 425 430
Elapsed time (hrs)
Elapsed time (hrs)

Figure 19: Well 2: Verification of deconvolution Figure 20: Well 2: Adapted rates

10000

2000 700
Rate Normalised nm(p) Change and Derivative (psi)

1000 FP 12 600

1000
500

FP 39
100

Oil Rate (STB/D)


Pressure (psia)

400

Radial flow stabilization 300

10
200
-1000

100
Rates from #(1-39)[1-39]{1.00000E+08}3770
1
0.001 0.01 0.1 1 10 100 -2000 0
Elapsed time (hrs) 13 14 15 16 17 18 19 20 21 22
Elapsed time (hrs)

Figure 21: Well 3: Adapted rate normalized log-log plot Figure 22: Well 3: Pressure and rate histories

DST with increasing liquid level during drawdowns


Fig. 22 shows the rate and pressure histories for a DST in oil Well 3. The pressure increases during the drawdowns, which
indicates a liquid level in the well. The rates are reported as constant during the drawdowns, which is incorrect. The usual
procedure is to convert pressures into fluid levels and use these to calculate the actual rates. Alternatively, rates can be
obtained by deconvolution.
To that end, the first and second drawdowns are divided into 4 and 12 different periods at constant rate, respectively, prior to
deconvolving the entire pressure history with rate adaptation. Results are shown in Fig. 23. For comparison, deconvolved
derivatives for the two build ups, FP7 and 21, and for the entire pressure history without rate adaptation are also included. The
default λ value for deconvolving the entire history is 3.68106 105 with and without rate adaptation. λ has to be increased to 106
with rate adaptation and to 3.68106 107 without rate adaptation to eliminate oscillations. The resulting deconvolved derivatives
are only slightly different. The pressure histories calculated from the deconvolved derivatives, with and without rate
adaptation, are compared to the actual pressure history in Fig. 24. With the adapted rates also shown in Fig. 4, the match is
good. Without rate adaptation, only the build ups are matched.
[SPE 134534] 8

10 2000 700

600

1000
500
Deconvolved Derivative

Oil Rate (STB/D)


Pressure (psia)
400

0 (1-21)[1-21]{3.68106E+07}1707.5 without rate adaptation


7
(1-21)[1-21]{1E+06}1707.5 with rate adaptation 300
21
0.1
77
#(1-21)[7]{9.34208E+04}1707.50
7
#(1-21)[21]{2.35605E+04}1707.50 200
-1000
#(1-21)[1-21]{1.00000E+06}1707.50 with rate adaptation

#(1-21)[1-21]{1.00000E+06}1707.50 without rate adaptation 100

#(1-21)[1-21]{3.68106E+07}1707.50 without rate adaptation


0.01
0.001 0.01 0.1 1 10 -2000 0
13 14 15 16 17 18 19 20 21 22
Elapsed Time hrs
Elapsed time (hrs)

Figure 22: Well 3: Deconvolved derivatives Figure 23: History pressure matches with and without rate
adaptation

Although unlikely in the particular case of Well 3 DST, the increase in drawdown pressure could also be attributed to a
decrease in skin effect. In the same way, deconvolution cannot distinguish an error in rate from a change in skin factor. It is up
to the interpreter to make this distinction.

Discussion and Conclusion


This paper has presented examples for two main benefits of deconvolution: (1) the increase of the radius of investigation,
which allows seeing boundaries and connectivities not visible in individual flow periods; and (2) the correction of erroneous
rates and the determination of missing rates. Both require applying deconvolution to entire pressure history sequences,
including build up and drawdown data.
Although deconvolution is theoretically only valid for linear systems, it can be applied to systems that are pseudo-linear,
such as with gas and multiphase flow. As any other interpretation method, such as straight lines and derivatives, it cannot be
used blindly, but requires knowledge and intuition from the user. In particular, the validity of the deconvolution must be
established, by verifying that the pressure history calculated from the deconvolved derivative can closely reproduce the actual
pressure measurements.

Nomenclature
D matrix in curvature measure y vector of adapted rate
E error measure z coefficient of the deconvolved derivative
FP flow period λ regularization parameter
g instantaneous source function ν rate error weight
k vector in curvature measure λdef default value of regularization parameter
m number of pressure points
νdef default value of rate error weight
N number of flow periods
∆p vector of pressure drop data
p vector of measured pressure
normalized pseudo-pressure  µ Z 
p
pi initial pressure, psia 2p
 2 p  ∫ µ Z dp
nm(p) psia
q vector of measured rate   p0 p 0
t time

References
1. Aluko, O.A. and Gringarten, A.C., 2009:” Well Test Dynamics in Rich Gas Condensate Reservoirs under Gas Injection,” paper SPE
121848 presented at the 2009 SPE EUROPEC/EAGE Annual Conference and Exhibition held in Amsterdam, The Netherlands, 8–11 June.
2. Amudo C., Turner, J., Frewin J.; Kgogo, T., and Gringarten A. C., 2006:”Integration of Well Test Deconvolution Analysis and Detailed
Reservoir Modelling in 3D Seismic Data Interpretation: A Case Study,” paper SPE 100250 presented at the SPE Europec/EAGE Annual
Conference and Exhibition, Vienna, Austria, 12–15 June
3. Earlougher, R.C. Jr., 1977. Advances in Well Test Analysis, Monograph Series. Richardson, Texas: SPE, 5.
4. Gringarten A. C., 2006:” From Straight lines to Deconvolution: the Evolution of the State of the art in Well Test Analysis,” paper SPE
102079, presented at the 2006 SPE Annual Technical Conference and Exhibition, San Antonio, Texas, U.S.A., 24–27 September 2006;
SPEREE (Feb. 2008) 11-1 pp. 41-62.
5. Gringarten, A. C., von Schroeter, T., Rolfsvaag, T., and Bruner, J., 2003:"Use of Downhole Pressure Gauge Data to Diagnose Production
Problems in a North Sea Horizontal Well," paper SPE 84470, presented at the 2003 SPE Annual Technical Conference and Exhibition,
Denver, CO, Oct. 5 – Oct. 8.
6. Ilk, D., Anderson, D. M., Valko, P. P., and Blasingame, T. A., 2006: “Analysis of Gas-Well Reservoir Performance Data Using B-Spline
Deconvolution,” paper SPE 100573 presented at the Gas Technology Symposium, Calgary, Alberta, Canada, 15-17 May, 2006.
7. Ilk, D., Valkó, P.P. and Blasingame, T.A., 2005:”Deconvolution of Variable Rate Reservoir Performance Data Using B-Splines,” paper
SPE 95571 presented at the 2005 SPE Annual Technical Conference and Exhibition, Dallas, Tx, 9-12 Oct. 2005.
[SPE 134534] 9

8. Levitan, M. M., 2003:"Practical Application of Pressure-Rate Deconvolution to Analysis of Real Well Tests", paper SPE 84290, presented
at the 2003 SPE Annual Technical Conference and Exhibition, Denver, CO, Oct. 5 – Oct. 8.
9. Levitan, M. M., Crawford, G. E., and Hardwick, A., 2004:” Practical Considerations for Pressure-Rate Deconvolution of Well Test Data,”
paper SPE 90680 presented at the SPE Annual Technical Conference and Exhibition held in Houston, Texas, U.S.A., 26–29 September.
10. Levitan, M.M. and Wilson, M.R., 2010:” Deconvolution of Pressure and Rate Data From Gas Reservoirs With Significant Pressure
Depletion,” paper SPE 134261 presented at the 2010 Annual Conference and Exhibition held in Florence, Italy, 19-22 September
11. McKinley, R.M., 1971:”Wellbore transmissibility from Afterflow-Dominated Pressure buildup Data,” J. Pet. Tech. (Jan 1971)
12. Meunier, D.F., Kabir, C.S., and Wittmann, M.J.,1987. Gas Well Test Analysis: Use of Normalized Pressure and Time Functions. SPEFE 2
(4): 629–636. SPE-13082-PA. DOI: 10.2118/13082-PA.
13. Moore, G., 1991: Crossing the Chasm, HarperCollins Publishers Limited ISBN 0-06-051712-3
14. Ramey, H.J., Jr., 1970:”Short-Time Well test Data Interpretation in the Presence of Skin Effect and Wellbore Storage,” J. Pet. Tech. (Jan
1970) 97.
15. von Schroeter, T., Hollaender, F., and Gringarten, A. C., 2001: “Deconvolution ofWell Test Data as a Nonlinear Total Least Squares
Problem,” paper SPE 71574 presented at the 2001 SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana, 30
September – 3 October; SPEJ (Dec. 2004) pp. 376-390.
16. von Schroeter, T., Hollaender, F., Gringarten, A., 2002:"Analysis of Well Test Data From Permanent Downhole Gauges by
Deconvolution," paper SPE 77688, presented at the 2002 SPE Annual Technical Conference and Exhibition, San Antonio, TX, Sept. 29 –
Oct. 2.
17. Whittle, T. and Gringarten A. C., 2008:” The Determination of Minimum Tested Volume from the Deconvolution of Well Test Pressure
Transients,” paper SPE 116239, presented at the 2008 SPE Annual Technical Conference and Exhibition, Denver, Co., U.S.A., 21–24
September.
18. Whittle, T., Jiang, H., Young, S. and Gringarten A. C., 2009:”Well Production Forecasting by Extrapolation of the Deconvolution of the
Well Test Pressure Transients,” paper SPE 122299 presented at the 2009 SPE EUROPEC/EAGE Annual Conference and Exhibition held
in Amsterdam, The Netherlands, 8–11 June.

You might also like