Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Deconvolution of Well-Test Data as a

Nonlinear Total Least-Squares Problem


T. von Schroeter, F. Hollaender, and A.C. Gringarten, SPE, Imperial College

Summary rate errors is rendered obsolete. The mathematical formulation is


an instance of what is known as a TLS problem in the numerical
We present a new time-domain method for the deconvolution of
analysis literature and as an Errors-In-Variables problem in statis-
well test data which is characterized by three novel features: (1)
tics. TLS has become a standard approach in parameter estimation
Instead of the rate-normalized pressure derivative itself, we esti-
problems, but its application to well-test analysis seems to be new.
mate its logarithm, which makes explicit sign constraints unnec-
3. Regularization is based on a measure of the total curvature of
essary; (2) the formulation accounts for errors in both rate and
the deconvolved pressure derivative, instead of its average slope,
pressure data, and thus amounts to a Total Least Squares (TLS)
as in an earlier approach15 and Paper I. Here, the motivation is that
problem; and (3) regularization is based on a measure of the over-
slopes provide important information about the flow regime and
all curvature of its graph. The resulting separable nonlinear TLS
should therefore be preserved as much as possible.
problem is solved using the Variable Projection algorithm. A com-
The paper is organized as follows: The first two sections are
prehensive error analysis is given. The paper also includes tests
introductory and give a summary of the deconvolution problem in
with a simulated example and an application to a large field example.
well-test analysis and a concise survey of its treatment in the
petroleum engineering and hydrology literature. Based on the
Introduction mathematical framework developed in these sections, we then give
a comprehensive account of our own approach. We also derive
With current trends towards permanent downhole instrumentation, analytic expressions for bias and variance of the estimated param-
continuous bottomhole well pressure monitoring is becoming the eter set based on simple Gaussian models for the measurement
norm in new field developments. The resulting well-test data sets, errors in pressure and rate signals. We illustrate our method with
recorded mainly during production, can consist of hundreds of a small simulated data set, demonstrating the effect of varying
flow periods and millions of pressure data points stretched over levels of regularization on the confidence intervals. The final section
thousands of hours of recording time. Such data sets contain in- presents an application to a large field example which allows a direct
formation about the reservoir at distances from the well which can comparison of our method with conventional derivative analysis.
be several orders of magnitude larger than the radius of investi-
gation of a single flow period.
Conventional derivative analysis is thus ill equipped to access Deconvolution and Well-Test Analysis
the full potential information content. What is required is an analy-
The primary quantity of interest in well-test analysis is the well-
sis method which can extract the response which the reservoir
bore pressure drop over time t for production at constant rate. More
would exhibit when subjected to a single drawdown at constant
specifically, we shall consider the rate-normalized wellbore pres-
rate over any period of time up to the entire production period. In
sure drop G(t) (i.e., the pressure drop at time t caused by constant
mathematical terms, this is a deconvolution problem. Since its first
production of one unit of rate measurement starting at time t⳱0).
formulation by Hutchinson and Sikora in 1959,1 it has received
sporadic, but recurring attention.2–17 This quantity has also been called the resistance function,1 the
This paper presents a new approach which is based on a regu- influence function,3 and the unit response function.4 If the flow in
larized, nonlinear TLS formulation. It is an update on earlier ver- the reservoir is governed by a set of equations which are linear in
sions which were presented at the SPE Annual Meetings in 200118 pressure and production rate, then the pressure drop ⌬P(t) ob-
and 200219 (henceforth referred to as “Paper I” and “Paper II”). served in a well test with time-varying production rate Q(t) satis-
More recently, our approach was taken up by Levitan,20 who sub- fies an integral equation of the form
t
jected it to a critical evaluation and suggested some modifications.
In terms of the usual classification into time-domain and spec-
tral approaches, ours is a time-domain approach. It differs from

⌬P共t兲 = 兵Q * g其共t兲 := Q共t⬘兲g共t − t⬘兲dt⬘. . . . . . . . . . . . . . . . . . . (1)
0
earlier approaches in this category in three important ways:
1. The solution is encoded in terms of the logarithm of the Here g denotes the impulse response, which is the ordinary time
rate-normalized pressure derivative, which automatically ensures
derivative of the rate-normalized wellbore pressure drop:
strict positivity of the derivative itself at the expense of rendering
the problem nonlinear. However, we are thus able to avoid explicit
constraints on the solution space which made previous constrained dG
approaches so difficult, yet still cannot prevent zeros in the de- g共t兲 = Ġ共t兲 := . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2)
dt
convolved derivative.
2. A new error measure accounts for uncertainties not only in
the pressure, but also in the rate data, which are usually much less An alternative version of the integral equation is
t
accurately known. Thus, provided sufficient data are available, our
method can provide a joint estimate of initial pressure, rates, and
response parameters; the time-consuming manual correction of
⌬P共t兲 = 兵Q̇ * g其共t兲 := 兰Q̇共t⬘兲G共t − t⬘兲dt⬘; . . . . . . . . . . . . . . . . . (3)
0

the two versions are mathematically equivalent if G(0)⳱0, which


follows from its definition, and Q(0)⳱0.
Copyright © 2004 Society of Petroleum Engineers
Eqs. 1 and 3 are known as Duhamel’s principle; they were
This paper (SPE 77688) was first presented at the 2002 SPE Annual Technical Conference originally discovered in the context of heat conduction and seem to
and Exhibition, San Antonio, Texas, 29 September–2 October, and revised for publication.
Original manuscript received for review 9 January 2003. Revised manuscript received 2
have been introduced into well-test analysis by van Everdingen
September 2004. Paper peer approved 10 November 2004. and Hurst.21 In situations in which nonlinearities play an important

December 2004 SPE Journal 375


role, as for instance in gas or multiphase flow, Duhamel’s principle Methods to accomplish this task fall into two categories: time-
is only approximately true over short time intervals.* domain methods and spectral methods.
Since the introduction of pressure-derivative analysis twenty
years ago,24 the focus of study has shifted to the derivative of the Time-Domain Methods. These attempt to solve the integral equa-
rate-normalized pressure drop with respect to the logarithm of time, tion (Eqs. 1 or 3) directly in the time domain. Even at the most
basic level, this involves a choice of interpolation schemes for the
dG共t兲 functions Q and g (respectively, Q̇ and G) in order to evaluate the
␥共t兲 := = tg共t兲. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (4) convolution integral. Usually, rates are modeled as stepwise con-
dlnt stant or piecewise linear, and the response as piecewise linear. The
resulting linear system has triangular structure and can therefore be
Providing a good estimate of this quantity from measurements solved explicitly in terms of a recursion relation. A number of
of pressure and production rate is therefore a task of crucial im- early approaches have exploited this structure1,2,4,5,7,8; some of
portance. Up to multiplication with time, this task amounts to them were developed specifically for the analysis of gas reser-
estimating the impulse response g from the linear relation (Eq. 1) voirs.2,5 However, the direct recursive approach is numerically
given ⌬P and Q as functions of time. In mathematical terms, this unstable, as already Hutchinson and Sikora1 observed. This is due
is a deconvolution problem. Common well-testing practice ignores to a combination of the following two factors:
this fact and uses estimation schemes which • The data are contaminated by measurement error; therefore,
• produce estimates for each flow period in isolation, it is not the measured signals ⌬P and Q that satisfy Duhamel’s
• involve numerical differentiation of the pressure signal, and principle, but instead the true, but unobserved, signals which differ
• attempt to account for the effect of earlier production by from ⌬P and Q by two error signals.
replacing ordinary time with a “superposition function.”25,26 The • The recursive nature of the algorithm causes these errors to
problem with this function is that it assumes a particular flow accumulate. It is true in general that discretizations of inverse
regime (e.g., radial flow, which leads to a systematic distortion of problems such as Eq. 1 lead to linear systems with highly ill-
the pressure derivative when flow is not radial). conditioned coefficient matrices.29,30 Thus, even if the solution is
This approach is fraught with numerical and conceptual diffi- unique, it tends to be extremely sensitive to errors in the data.
culties; moreover, it can only yield estimates of the response up to This numerical instability usually manifests itself in large os-
the maximal elapsed time during each flow period. In practice, the cillations of the deconvolved response in the presence of even very
radius of investigation is even further reduced in the process of modest levels of data error. Here, the sensitivity to rate errors
numerical differentiation, which amplifies measurement errors just appears to be far more dramatic; for an analytic discussion, see
as it amplifies reservoir features.27 Kuchuk et al.9,13
By contrast, deconvolution estimates are by definition esti- The instability problem has prompted various responses:
mates for varying production rate, and are therefore not subject to • Kučuk6 and Thompson and others7,8 investigated a number
any constraints on the radius of investigation short of the test of refinements to the interpolation schemes.
duration.** Besides, they are based entirely on the linear system • Coats et al.3 were the first to reformulate deconvolution as an
(Eq. 1) and therefore avoid numerical differentiation as well as any optimization problem; recent developments have followed their
bias due to assumptions about the reservoir model. Thus the es- lead.10,13,15 Here, the idea is to reduce the number of parameters
sence of deconvolution vis-a-vis conventional analysis could be in the solution such that the discretized linear system becomes
summarized as “estimating the derivative without taking it.” overdetermined, and to seek solutions by minimizing a measure
Beyond these common principles, a considerable variety of of the error signal(s). A family of such measures is given by the
deconvolution approaches has emerged in the literature; a brief Ln norms,

再兰 冎
survey will be given in the following section. However, so far none
T 1Ⲑn
of these approaches appears to have met with universal success; in 1
particular, the ones that have been tested with simulated signals 㛳u㛳n := |u共t兲|ndt ,1 ⱕ n ⬍ ⬁. . . . . . . . . . . . . . . . . . . . . (5)
seem to share a disproportionate sensitivity to measurement inac- T
0
curacies, with authors reporting uninterpretable results in the pres-
ence of error levels as modest as 1 to 2% in the rates. This may The norms used so far are the ones for n⳱1 and n⳱2, which
explain why deconvolution has not yet become the standard tool in lead respectively to Linear Programming and to a Least Squares
well-test analysis which it ought to be. (LS) formulation. (The error measure used by Gajdica et al.10 is a
weighted version of the L1-norm with weight function w; their
formulation includes the one due to Coats et al. as the special
Previous Work case w≡1.)
For the sake of simplicity, the following assumptions will be made • A number of authors2,3,10,13,15 implemented constraints on
throughout this section: the solution space. Hutchinson and Sikora1 argued on physical
• The initial pressure p0 is known.† grounds that the influence function should not only be positive and
• The signals ⌬P and Q have been measured over a time in- increasing, but also concave; Katz et al.2 and Coats et al.3 made
terval [0,T], where T>0 is the test duration and Q≡0 outside this this statement more precisely:
interval. (Thus, it does not matter whether the upper limit of in-
tegration Eqs. 1 and 3 is t, T, or ⬁.) G ⱖ 0, Ġ ⱖ 0, G̈ ⱕ 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (6)
It follows from Duhamel’s principle that only the part of g and For single-phase, slightly compressible Darcy flow with initial
G over the interval [0,T] has any bearing on the data, and therefore equilibrium, these constraints were derived rigorously by Coats et
that this part of the response signals is all one can hope to recover. al.,3 who showed that in this case there are sign constraints for
derivatives of any order. The method by Kuchuk et al.13 uses the
same constraints, but measures the pressure error in L2-norm.
* For flow problems involving a gas phase, linear approximations of the governing equa- Baygün et al.15 use a different set of constraints which they ex-
tions have been developed in terms of various combinations of pressure, compressibility, press in terms of the logarithmic derivative ␥, Eq. 4; they constrain
and real gas factors; these are known as (single22 or twophase23) pseudopressures. .
Duhamel’s principle is then approximately satisfied in the form shown in Eq. 1 if the its “energy” 㛳␥㛳22 and the values of its autocorrelation function
pseudopressure drop is substituted for ⌬P.
** There are of course situations in which it may be desirable to consider only a small portion 共␥,␥s兲
of the pressure sequence, for instance to test for changes in reservoir behavior28 (and ␳共s兲 = . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (7)
thus for the extent to which Eq. 1 is satisfied for the data set in question). However, the 㛳␥㛳2㛳␥s㛳2
important point is that with deconvolution, the size of the pressure data window is entirely

a matter of choice.
To our knowledge, the only other method besides our own to provide an estimate for p0
for shifts s⳱m⌬t, m⳱1, 2, where ⌬t is a fixed timestep size
is the one due to Baygün et al.15 between samples, ␥s the time-shifted signal t哫␥(s+t) and

376 December 2004 SPE Journal


T spectral representations of deterministic as well as stochastic pro-

1
共u,␷兲 := u共t兲␷共t兲dt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (8) cesses, in particular to the following:
T • The continuous Laplace and Fourier transforms.
0
• Their discrete counterparts, such as the discrete Fourier trans-
the scalar product of two signals. The autocorrelation function form and the z transform.
takes values in the interval [–1,1]. Therefore, constraints of the • The associated magnitude and power spectra.
form ␳(m⌬t)ⱖ␪m for ␪m>–1 enforce a degree of correlation be- • The spectral density function of a stationary stochastic pro-
tween a function and its time-shifted copies. cess.31
A synopsis of approaches is given in Table 1. We have focused The convolution theorem states that the transform of a convo-
on the basic mathematical structure of the formulations and ne- lution product is equal to the product of transforms. In symbols:
glected technical details such as smoothing (which was considered
by Hutchinson and Sikora1) and interpolation issues (which are
⌬P = Q * g ⇔ ⌬P = Q ⭈ g, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (9)
discussed at length in the papers by Thompson et al.7,8).
To what extent these methods succeed is often hard to establish
from the publications alone. Early developments were usually where the bar denotes the spectral transform in question. Thus, in
aimed at estimating specific quantities such as oil in place, or at principle, the estimation problem reduces to the following recipe:
predicting productivity, and hence any attempts at verification • From the data ⌬P and Q, estimate the spectral transforms ⌬P
were entirely based on these derived data. Only later did the accuracy and Q;
of the response obtained from simulated signals become the dominant • Hence, obtain an estimate of the transformed response as
criterion, and only the papers by Kuchuk et al.13 and Baygün et al.15 g⳱⌬P/Q; and
reported tests with simulated signals perturbed by measured amounts • If the transform is invertible, apply the inverse transform to
of random errors. Their conclusions are as follows: g to obtain the response estimate g, or interpret the results in the
• Irrespective of the interpolation scheme, results obtained spectral domain (as was suggested by Bourgeois and Horne,14 who
with the linear recursion method show strong oscillations if the also developed a catalogue of type curves in the Laplace domain).
rate errors reach levels of between 0.1 and 1%. The transform commonly used in well-test analysis is the
• The constrained Least Squares approach proposed by Ku- Laplace transform. Estimates of the transforms ⌬P and Q require
chuk et al.13 is reliable up to 2% rate error. interpolation schemes for the signals ⌬P and Q; various schemes
• Deconvolution with energy and autocorrelation constraints have been considered in the literature.11,12,16 Once the interpola-
(Baygün et al.15) was developed and tested for pressure-pressure tion schemes are chosen, the response transform estimate g is
deconvolution; therefore, both input signals have similar error lev- defined for all values of the spectral variable, and can be trans-
els. The method proved reliable up to error levels of 1%; from 2% formed back to the time domain using Stehfest’s efficient inver-
error, the deconvolved response shows oscillations about the true sion algorithm.32
response at late times. The spectral approach is attractive for its simplicity and direct-
With current technology, error levels of 1% and below are only ness, at least at the algorithmic level. The apparent simplicity may
achievable for pressure measurements; rates are usually in error by be to blame for a shortage of systematic error analysis to date; only
at least 5 to 10% even if they are continuously measured. It follows Onur and Reynolds16 reported tests in which simulated signals
that none of the approaches discussed previously is likely to work were subjected to random perturbations. It is beyond the scope of
with realistic rate and pressure data; however, the constrained our paper to address this gap; instead, we shall only mention a
Least Squares methods appear to be robust enough for pressure- serious limitation which has been identified in the context of the
pressure deconvolution.15 Laplace transform, and which we shall refer to as the missing data
problem. As the problem arises from basic mathematical properties
Spectral Methods. These11,12,14,16,17are based on the convolution common to all spectral transforms, it is likely to affect other spec-
theorem of spectral analysis. This theorem applies to a number of tral approaches as well.

December 2004 SPE Journal 377


With data in the time interval [0,T], one can at best estimate The implicit alternative, and the only way to impose strict
inequality constraints, is to parametrize the solution space in such
T



a way that the constraints are automatically satisfied. The obvious
⌬P(s)⳱ ⌬P(t)e-stdt, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (10) candidate for such an encoding is the one which is used in the
0 diagnostic plot. To avoid a profusion of numerical factors, we shall
use natural instead of decadic logarithms; thus, we propose to
whereas the convolution theorem (Eq. 9) holds strictly only for estimate
Laplace transforms of functions defined on the entire positive time
axis (i.e., for T⳱⬁). Thus, even if pressure data in the test interval Z共␶兲 = ln兵tg共t兲其 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (11)
are complete and the initial pressure p0 is known, the Laplace as a function of ␶⳱lnt. In terms of this encoding, Duhamel’s
transform of the pressure drop can only be correctly estimated if, principle becomes
at the end of the test, the pressure has returned to its equilibrium t lnt

兰 兰Q共t − e 兲e
value. Whether this happens at all depends on the characteristics of dt⬘ ␶ Z共␶兲
⌬P共t兲 = Q共t − t⬘兲t⬘g共t⬘兲 = d␶. . . . . . . . . (12)
the reservoir, and even if it does, it would be uneconomical for t⬘
0 −⬁
most reservoirs to interrupt production and wait until it happens.
Extrapolation methods have been investigated by various au- Note that the right side is still linear in the rate signal, Q, but
thors,11,14 but they require assumptions about the late-time reser- nonlinear in the response function Z. From a practical point of
voir behavior which, in the words of one of these papers,14 render view, the nonlinearity is of course unwelcome; however, on bal-
at least part of the calculated Laplace transform “nothing more ance, this slight complication is more than outweighed by the
than a self-fulfilling prophecy”; moreover, other authors have benefits.
found them “inappropriate when working with noisy data.”16
In any case, it is clear from the comments following Eq. 3 that 2. Total Least Squares Error Model. The more recent time do-
the response estimate with or without extrapolation must cease to main methods tend to be based on a Least Squares formulation,
be meaningful at the end of the test interval; according to Bour- and thus at least implicitly on an error model. However, the error
geois and Horne,14 the loss due to the missing data for t > T can model associated with an ordinary Least Squares approach attrib-
be up to half a log cycle in the response. Numerical experiments utes the entire difference between the two sides of Eq. 1 to errors
which we carried out ourselves suggest that the estimates g ob- in the pressure measurement. This is in sharp contrast with current
tained with the incomplete pressure signal up to the finite test well-testing practice, in which reported rates are often merely al-
duration T (and assumed zero for t>T) tend to follow the Laplace located; but even if they are measured, errors of up to 10% are not
transform of the correct response function truncated at time T (i.e., uncommon. In any case, the relative uncertainty in the rate signal
g̃ in the notation of Eq. 10). It is not difficult to quantify this is typically much larger than that in the pressure signal. If rate data
observation; its practical exploitation would require reliable nu- are to be used at all in the deconvolution process, the error model
merical methods to recover such truncated signals from their should certainly reflect the relative size of their contribution to the
Laplace transforms. However, this appears to be a notoriously overall error. This line of reasoning leads directly to a TLS for-
difficult problem itself, and the only algorithm we found in the mulation (see, for instance, Golub and van Loan,34 section 12.3,
literature33 did not seem accurate enough for our purpose. and Björck,35 section 4.6).
Another missing data problem which frequently occurs in prac-
tice is that parts of the pressure signal are not recorded. This would 3. Regularization Based on Curvature. Previous studies demon-
have little effect on a time-domain estimate unless the number of strated the need for regularization to maintain a degree of smooth-
recorded pressure points was insufficient to constrain the response ness in the solution in the presence of measurement errors; our
parameters. By contrast, the effect of gaps in the signals on a spectral method is no exception. We initially followed these studies in
estimate is much more dramatic as spectral transforms are by defini- experimenting with a regularization scheme based on the sum of
tion integrals over the entire time domain. However, for this situation, squared derivatives between nodes,18 but came to the conclusion
a more promising recovery procedure was suggested by Gilly and that penalizing instead a measure of the overall curvature of the
Horne.17 They treat the missing pressure data as additional param- response graph yields superior results.19 Here, the motivation is
eters in a regression procedure based on the difference ⌬P−Q⭈g. that the slopes of the diagnostic plot provide crucial information
about the flow regimes and should therefore be preserved as much
Deconvolution of Well-Test Data as a Nonlinear as possible.
Total Least Squares Problem None of these three ingredients is sufficient on its own; for
instance, experiments with an unregularized, unconstrained linear
Our own approach to deconvolution is a time-domain approach; it TLS approach still led to oscillations and even negative values in
is characterized by three novel ingredients which are motivated by the response when rates were contaminated with errors of 10%. A
the results of earlier studies and our own experiments: summary of this approach appeared in Paper I18 and will not be
repeated here.
1. Implicit Constraint Encoding. Experience has shown that un- In order to present a mathematical formulation which incorpo-
constrained deconvolution methods are numerically unstable. Thus rates all three principles, we shall denote the wellbore pressure and
it is necessary to restrict the space of admissible solutions by pressure drop measured at time ti, i⳱1…m by pi⳱P(ti) and
imposing physically reasonable constraints. The most obvious con- ⌬pi⳱p0−pi, respectively. We also assume that interpolation
straint is that the solution should be positive. schemes for rate and response have been chosen such that
There are two fundamentally different ways to implement con- N N

兺q ␪ 共t兲, Z共␶兲 = 兺z ␺ 共␶兲. . . . . . . . . . . . . . . . . . . . . . (13)


straints, namely explicitly and implicitly. The explicit approach
imposes constraints in the form of additional equations or inequal- Q共t兲 = j j k k
j=1 j=1
ities to be satisfied by the solution, or its parameters. This is the
approach that all constrained formulations to date have taken. In In the rate interpolation, the coefficients qj are known from mea-
practical terms, explicit constraints add a substantial degree of surements, whereas in the interpolation of the pressure derivative
complication to the algorithm. In mathematical terms, what is at response, the coefficients zk are to be determined in the deconvo-
issue is the topology of the solution space: The only kind of in- lution process. Typically, flow rates are reported as stepwise con-
equality that can be enforced by explicit constraints is the nonstrict stant over time; in this case, qj is the production rate over a time
kind, g(t)ⱖ0; however, only the strict inequality g(t)>0 makes interval ajⱕtⱕbj, and
physical sense, at least for t>0. Indeed, the conventions for the
diagnostic plot are such that the logarithms log 10 G(t) and

1 if aj ⱕ t ⱕ bj,
␪j共t兲 = 0 otherwise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (14)
log10{tg(t)}⳱log10{dG/dlnt} are graphed over log10t, which Now evaluating Eq. 1 at times t⳱ti yields a set of equations which,
moves the boundary g≡0 to negative infinity! in matrix-vector form, can be written as

378 December 2004 SPE Journal


⌬p = C共z兲q, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (15) as x⬘⳱[p0,y]T and x⬙⳱z, we can write the error measure in a more
compact form as
where q⳱(q1,…,qN) and C is a matrix-valued function of the
response coefficients with components E共x⬘,x⬙兲 = 㛳F共x⬙兲x⬘ − ␷共x⬙兲㛳22, . . . . . . . . . . . . . . . . . . . . . . . . . . . (26)
lnT
where the matrix F(x⬙) and the vector ␷(x⬙) are given by
Cij共z兲 = 兰␪ 共t − e 兲e
j i
␶ Z共␶兲
d␶ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (16)
⌼m −C共x⬙兲

冤 冥
−⬁

for i⳱1…m and j⳱1…N. The case we shall consider in detail is F共x⬙兲 = 0 公␯ IN , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (27)
that of stepwise rate functions and linear interpolation of the re- 0 0
sponse: We choose nodes ␶k such that

冤 冥
p
−⬁ = ␶0 ⬍ ␶1 ⬍ ␶2 ⬍ . . . ⬍ ␶n = lnT . . . . . . . . . . . . . . . . . . . . (17)
and ␷共x⬙兲 = 公␯ q . . . . . . . . . . . . . . . . . . . . . . . . . . . . (28)
and interpolate Z linearly between values zk⳱Z(␶k) which are to be
computed: 公␭共Dx⬙ − k兲
Z共␶兲 = ␣k + ␶␤k of ␶k−1 ⱕ ␶ ⱕ ␶k, . . . . . . . . . . . . . . . . . . . . . . . (18) Here, IN denotes the identity matrix of size N. Note that the residue

where we assume ␤1⳱1 to model wellbore storage before the first r共x⬘,x⬙兲 = F共x⬙兲x⬘ − ␷共x⬙兲 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (29)
node (which must therefore be chosen early enough), and depends linearly on x⬘ and nonlinearly on x⬙. Thus, in mathemati-
zk − zk−1 cal terms, minimizing the full error measure amounts to a sepa-
␤k = , ␣ = z − ␤k␶k. . . . . . . . . . . . . . . . . . . . . . . . . . . . (19) rable nonlinear Least Squares problem.35
␶k − ␶k−1 k k This formulation covers two estimation problems which de-
For this case, expressions for the coefficients Cij(z) and their de- serve separate mention. They differ only in the choice of variables
rivatives are given in Appendix A. For a more general formulation taking part in the minimization process:
of Duhamel’s principle which is based on the same response en-
coding but does not require any assumption about the behavior 1. Response Parameters Only. Here, the initial pressure and rates
before the first node, see Levitan.20 are kept fixed, and thus the second term in Eq. 23 does not arise.
In the presence of measurement uncertainty, both observed The response coefficients z are determined by minimizing
signals ⌬p and q contain errors. Let us denote the true, but unob-
served, signals by ⌬p+␧ and q+␦, where ␧ and ␦ are signals rep- E1共z兲 = 㛳⌬p − C共z兲q㛳22 + ␭㛳Dz − k㛳22, . . . . . . . . . . . . . . . . . . . . . (30)
resenting the measurement errors in pressure and rate, respec- which is a nonlinear Least Squares problem.
tively. These unobserved signals satisfy Duhamel’s principle:
⌬p + ␧ = C共z兲共q + ␦兲. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (20) 2. Initial Pressure and Rate Correction. Here, the response is
assumed known, and the true rates and initial pressure are to be
The TLS formulation of the deconvolution problem is to find the estimated from measured pressure and rate data. Thus the last term
response coefficients z with the “smallest” perturbations ␧ and ␦ in Eq. 23 is constant and can be omitted; moreover, C=C(z) is a
and the least overall curvature such that Eq. 20 holds. If we mea- constant matrix, and so the relevant error measure is
sure the size of the perturbations in ᐉ2−norm*
E2共p0,y兲 = 㛳p0⌼m − p − Czy㛳22 + ␯㛳y − q㛳22. . . . . . . . . . . . . . . . . (31)
㛳x㛳2 := 公x x = 公 + T
x21 x22 + . . ., . . . . . . . . . . . . . . . . . . . . . . (22)
Here, the residue is linear in the variables, and thus for given
we are led to consider a class of error measures of the form weight ␯ the estimation problem amounts to a linear Least Squares
problem. There is a unique solution (p0,y) with minimal vector
E = 㛳␧㛳22 + ␯㛳␦㛳22 + ␭␬共z兲2, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (23) 2-norm.35
where ␯ and ␭ are fixed weights and ␬(z) is a measure of the At first glance, it might seem that any estimate involving both
curvature of the graph of Z, taken as a function of its coefficients. rates and responses must necessarily be underdetermined, as Du-
It is shown in Appendix B that an approximate measure of the hamel’s principle is invariant under a rescaling of the form
curvature can be written in the form
Q 哫 ␣Q, g 哫 ␣−1g . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (32)
␬共z兲 = 㛳Dz − k㛳2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (24)
for any positive ␣ (cf., Eq. 1). In fact, unless ␯⳱0 (i.e., all rate
with a constant matrix D and a constant vector k. In terms of these information is ignored), the error measure E does not share this
quantities, we define our estimate for given weights (␯,␭) as the scaling ambiguity: Closer inspection shows that the corresponding
global minimizer** of the error measure (Eq. 23) subject to the transformation of estimated rates and response parameters, viz.
constraint (Eq. 20). Alternatively, expressed in terms of z, the
initial pressure p0, and the true, unobserved rates y=q+␦, our es- y 哫 ␣y, z 哫 z − ⌼nln␣, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (33)
timate is the global minimizer of leaves only the pressure match and the regularization term invari-
E = 㛳p0⌼m − p − C共z兲y㛳22 + ␯㛳y − q㛳22 + ␭㛳Dz − k㛳22, . . . . . . . . (25) ant, but not the rate match. Thus, among all multiples of the
optimal rate vector y, the error measure selects the one with mini-
where ⌼m is the vector of dimension m with each component equal mal Euclidean distance from the given rates q. (Nonetheless, the
to 1. We refer to 㛳 ␦ 㛳 2 ⳱㛳y−q㛳 2 as the rate match and to joint estimate can of course be underdetermined; for instance, this
㛳␧㛳2⳱㛳p0⌼m−p−C(z)y㛳2 as the pressure match. Grouping variables is certainly the case if m<n.)

Error Analysis and Regularization


* An alternative, and perhaps preferable, way to quantify the rate error would be to think of The purpose of this section is to derive statistical properties of the
it as a continuous signal and to measure its size in L2 norm:
deconvolution estimate, and to discuss criteria for optimal choices
N N
of the two parameters ␯ and ␭. Under certain assumptions, these
⌬共t兲 = 兺␦ ␪ 共t兲, 㛳␦㛳 = 兺␦ ␦ 共␪ ,␪ 兲. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (21)
j=1
j j
2
2
j,k=1
j k j k
issues turn out to be closely related. Whether these assumptions are
For the stepwise interpolation scheme of Eq. 14, this would amount to replacing 㛳␦㛳22 in met in practice depends on the data quality. Thus we also give
Eq. 23 by a weighted sum of squared rate errors, where the weights are the durations of
the flow periods. more robust, yet also more subjective criteria, which, in our
** It is beyond the scope of this paper to address the question of uniqueness of the solution. experience, still work when the statistically based criteria fail; this
At a local level, the use of the generalized matrix inverse in the joint linearized estimate
(discussed later) ensures that in the case of rank deficiency of the matrix A, the solution was the case with the rate data in the field example discussed later
with minimal vector 2-norm is chosen; this solution is unique. Cf. Björck.35 in this paper. For the sake of completeness, we present the error

December 2004 SPE Journal 379


冤 冥
analysis for the joint estimate of initial pressure, rates, and pressure ␴2pIm 0 0
derivative. In Paper II,19 where this analysis was first presented,
the initial pressure was assumed known; hence, there are slight cov兵b其 = 0 ␯␴2qIN 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . (43)
differences in some of the expressions. 0 0 0
We shall assume that perturbations in the observed quantities p
Here, the entire matrix is square and has size m+N+d, where d is
and q are sufficiently small to allow linearization of the residue
the row dimension of the regularization matrix D. Only the diago-
about the true rate and response parameters, y* and z*. To do this,
nal elements ␴2i ,i⳱1. . .N+n+1 of cov{x̂} are required for the
we shall write
confidence intervals; taking 2␴i as their radii,* we can write the
C共z兲y ≈ C*y + ⌫共z − z*兲, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (34) resulting confidence region in vector form as
R = 关x* + ⺒兵x̂其 − 2␴, x* + ⺒兵x̂其 + 2␴兴. . . . . . . . . . . . . . . . . . . (44)
where C* and ⌫ are constant matrices given by
These regions are not only useful for forward simulation, but
C* = C共z*兲, ⌫ = ⵜz兵C共z兲y*其z=z . . . . . . . . . . . . . . . . . . . . . . . . . (35) also for data analysis, because x̂∈R is equivalent to
*

With the substitution in Eq. 34, we can approximate the error x* ∈关x̂ − ⺒兵x̂其 − 2␴, x̂ − ⺒兵x̂其 + 2␴兴. . . . . . . . . . . . . . . . . . . . . (45)
measure (Eq. 25) in the form Thus, provided that the deconvolution estimate and the true
E共x兲 ≈ 㛳Ax − b㛳22, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (36) solution are close enough for their bias and variance to be approxi-
mately equal, Eq. 45 can be taken as an estimate for the confidence
where x⳱(p0,y,z)T and region of the true solution.
If this type of error analysis is to be applied to field data,
⌼m −C* p − ⌫z*

冤 冥 冤 冥
−⌫ estimates of the error variances ␴2p and ␴2q are required. At present
A= 0 公␯IN 0 ,b= 公␯q . we can only offer a posteriori estimates (i.e., estimates that depend
on the deconvolution estimates, and thus in turn on the choices of
0 0 公␭D 公␭k the two weights ␯ and ␭).
An unbiased estimate of the pressure variance is given by
This is a standard Least Squares problem, and thus its error analy-
sis is straightforward. The estimate is the unique minimizer of Eq. m


1
36 with minimal vector 2-norm, which is given by ␴ˆ 2p = c2, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (46)
m − N i=N+1 i
x̂ = A b, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (37)

where c⳱UT⌬p and C*⳱UDC*VT is the singular value decompo-

where A denotes the generalized (Moore-Penrose) matrix in- sition34 of the matrix C* (cf., Björck,35 Theorem 1.1.1). In our
verse35 of A. In the generic case in which A has full column rank, experience, this works well with simulated data if z* is taken as the
response estimate.
A† = 共ATA兲−1 AT, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (38) For the rate variance ␴2q, the answer depends on the statistical
assumptions; in the most favorable case, this answer also provides
which is the form familiar from the normal equations.* a criterion for the optimal weight parameter ␯. A sufficient set of
Bias is defined as the difference between expected and true assumptions is that, in addition to Eqs. 39 and 41, the components
value of the estimate, where the expected value can be understood of ␧ and ␦ are normally distributed; in symbols:
as an average over a large number of realizations of the random
variables. This is where assumptions about the error sequences ␧ ␧ ∼ N共0,␴2pIm兲, ␦ ∼ N共0,␴2qIN兲. . . . . . . . . . . . . . . . . . . . . . . . . . . (47)
and ␦ come into play. For the derivation of the bias, we assume
that each component of these sequences is a random variable with In this situation, the following statements hold:
zero mean: • The Least Squares estimate of rates and initial pressure co-
incides with the maximum likelihood estimate if
⺕兵␧其 = 0, ⺕兵␦其 = 0, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (39)
␯ = ␴2p Ⲑ ␴2q. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (48)
where ⺕ {} denotes expectation. It is shown in Appendix C that
with this assumption, and provided the matrix A has full rank, the (Indeed, the maximum likelihood estimate minimizes20,36
bias of the parameter estimate (Eq. 37) is given by 㛳p0⌼m − p − Cy㛳22 㛳y − q㛳22
␹2共p0,y兲 = + . . . . . . . . . . . . . . . . . . . (49)

冤 冥
0 ␴2p ␴2q
⺒兵x̂其 ≡ ⺕兵x̂其 − x* = ␭共ATA兲−1 0 . . . . . . . . . . . . (40) which is a constant multiple of E2(p0,y) (Eq. 31) if and only if Eq.
DT共k − Dz*兲 48 holds.)
• Moreover, this value of ␷ minimizes the Generalized Cross-
This shows that, at least in the generic case (ATA invertible, Validation (GCV) error; see Golub et al.,37 Theorem 2 (where
DT(k−Dz* )⫽0), the estimate is biased unless ␭⳱0. n,␭,␴,␣, and ␤ correspond, respectively, to N,v/N,␴p,␴2q and ␦ in
In order to derive confidence intervals, we need to make the our notation).
error model more specific. We shall now further assume that the Thus, minimization of the GCV error provides both the optimal
perturbations satisfy error weight ␯GCV and an estimate for the rate variance:

⺕共␧␧T,␦␦T,␧␦T兲 = 共␴2pIm,␴2qIN,0兲. . . . . . . . . . . . . . . . . . . . . . . . . . (41) ␴ˆ 2q = ␴ˆ 2p Ⲑ ␯GCV. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (50)

This condition means that the components of ␧ and ␦ are uncor- When used with simulated data designed according to Eq. 47,
related random variables with variances ␴2p and ␴2q, respectively. and with C evaluated at the true response z*, minimization of the
By definition, the covariance matrix of the estimate is GCV error and Eq. 50 provided reasonable a posteriori estimates
of the rate variance. This suggests that GCV may hold out some
cov兵x̂其 = ⺕兵共x̂ − ⺕兵x̂其兲共x̂ − ⺕兵x̂其兲T其 = A†cov兵x̂其 A†T; . . . . . . . . . (42) prospect at least for iterative adaptation of ␯ with field data,
provided the rates are genuinely measured and not merely allocated.
the covariance of the vector b is given in matrix block form by

* This corresponds to the 95% confidence interval of a single parameter estimate if the
* Note, however, that these formulas are not recommended for computation! The standard measurement errors are normally distributed. (For the probabilistic interpretation of con-
way to compute x̂ involves the Singular Value Decomposition of the matrix A; see Björck.35 fidence regions, see Press et al.,36 section 14.5.)

380 December 2004 SPE Journal


Unfortunately, in most field examples we have so far had ac- given rates are non-negative (i.e., there is no injection); otherwise,
cess to, the rates were allocated; with these examples GCV mini- to the first pressure data point.
mization either failed or led to such low values of ␯ as would As initial response we take a unit slope before the first node ␶1
amount to ignoring the rate data, with the risk of an underdeter- and a constant value thereafter


mined estimate. We took this as an indication that the Gaussian
ln c + ␶ if ␶ ⱕ ␶1,
error model was inappropriate for these data. Lacking better esti- Z0共␶兲 = ln c0 if ␶ ⱖ ␶1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . (54)
1
mators, we used the mean squared rate match,
Here, continuity at the first node implies
␴ˆ 2q = 㛳y − q㛳22 Ⲑ N. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (51)
ln c0 + ␶1 = ln c1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (55)
In simulated examples, this estimator did not perform much worse
The value of c1 is determined such that the best pressure match is
than the GCV-based estimator; both underestimated the true rate
obtained. In detail, evaluating Eqs. 1, 11, and 13 with the model in
variance by about 10 to 20%.
Eq. 54 leads to
The basic principle of GCV is that the estimate should not
overly depend on any single data point. For the joint estimate of N

rates, response, and initial pressure, this principle ideally ought to


be applied to the joint selection of both parameters ␯ and ␭. How-
⌬p共t兲 = c0⌬p1共t兲 + c1 兺q ␮ 共t兲,
j=1
j j . . . . . . . . . . . . . . . . . . . . . . . . (56)

ever, to our knowledge there is not even an algorithm for the where
linearized version of this problem (i.e., for the regularized linear lnT

兰␪ 共t − e 兲d␶ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (57)
Total Least Squares problem). ␶
With regard to the second parameter ␭, we briefly experi- ␮j共t兲 = j
mented with another criterion found in the statistical literature, ␶1
namely the “L curve” criterion.38 Our experience is that it consis- and
tently overestimates the optimal level of regularization by several
orders of magnitude, with the effect that important features of the |⌬p1共t兲|ⱕ exp共␶1兲 ⭈ max兵|qj|, j = 1 . . . N其. . . . . . . . . . . . . . . . . (58)
solution are lost. By choosing the first node early enough, the contribution ⌬p1 from
However, at least the effect of increasing ␭ is easy to observe the early part can be made arbitrarily small. Thus, we only evaluate
since it relates to the smoothness of the result. Thus, a possible the remaining part at the times ti and match the result to the giv-
way of choosing it is simply by looking at the result and increasing en pressure measurements. In vector notation, the mismatch is
␭ to a value for which the response is just smooth enough to be given by
interpretable without losing its dominant features. This is admit-
tedly a very subjective criterion, but, at the time of writing, it is the ␧ = c1Mq − ⌬p, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (59)
only one we can offer. where the components ⌬pi=p0−pi are computed with the starting
As default values we use value of p0 and M denotes the matrix
␮1共t1兲 ␮2共t1兲 ⭈⭈⭈ ␮N共t1兲

冤 冥
N 㛳⌬p㛳22 㛳⌬p㛳22
␯def = , ␭def = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . (52) ␮1共t2兲 ␮2共t2兲 ⭈⭈⭈ ␮N共t2兲
m 㛳q㛳22 m
M= · · · . . . . . . . . . . . . . . . . . . . . . (60)
where ⌬p is formed with the first, or maximal, pressure sample in · · ·
· · ·
place of the initial pressure. These values can be motivated by ␮1共tm兲 ␮2共tm兲 ⭈⭈⭈ ␮N共tm兲
noting that 㛳⌬p㛳22/m and 㛳q㛳22/N are, respectively, the average
Minimal mismatch 㛳␧㛳2 is obtained for
squared pressure drop and the average squared rate, and thus the
sum of relative squared pressure and rate errors, plus the curvature ⌬pTMq
term, equals c1 = . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (61)
㛳Mq㛳22
㛳⌬p − C共z兲y㛳22 㛳y − q㛳22 Our practical experience so far suggests that the method is stable
+ + 㛳Dz − k㛳22, . . . . . . . . . . . . . . . . . . . . (53)
㛳⌬p㛳22 Ⲑ m 㛳q㛳22 Ⲑ N in the sense that results do not depend on the starting point of the
iteration. The choice we just described has the merit of being
which is equal to mE/㛳⌬p㛳22 with ␯ and ␭ given by Eq. 52. If D and relatively easy to compute and not too distant from interpretable
k are scaled with the Frobenius norm 㛳D㛳F (see Appendix B), ␭def response curves, which ought to reduce the number of iterations
usually underestimates the required level of regularization by one required for convergence.
to three orders of magnitude; thus. we usually increase ␭ from ␭def
by successive powers of 10, leaving ␯ at its default value. Stopping Criterion. We monitor the decline ⌬E in the error mea-
sure (Eq. 25) from its initial value E0 and stop the iteration once
Nodes, Initialization, and Convergence the following convergence criterion is satisfied:␽
Three choices remain to be discussed in order to complete the
specification of the deconvolution estimate, namely: (1) the nodes ⌬E ⬍ ␽E0, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (62)
␶k, (Eq. 17); (2) starting values for rates, initial pressure, and where ␽ is a user-defined threshold which is normally in the range
response parameters; and (3) a stopping (or convergence) criterion. 10–8…10–5.
Here the nodes are always required, the other two only if response
parameters are part of the estimate. For if only rates and initial Nodes. From the preceding discussion and in particular from Du-
pressure are to be estimated, the residue is linear in the parameters, hamel’s principle, it follows that the choice of nodes should satisfy
and the estimate can be computed in a single step by solving a the following criteria:
linear Least Squares problem. If, however, response parameters are 1. The last node must be at ␶n⳱lnT, where T is the test duration
to be estimated as well, the residue is nonlinear, and the decon- (i.e., the time difference between the last pressure sample and the
volution problem amounts to a nonlinear or separable nonlinear start of the earliest drawdown period in the rate signal). In sym-
Least Squares problem. Algorithms for this type of problem are bols: T⳱tm–a1.
necessarily iterative and thus require starting values and a stopping 2. The first node ␶1 ought to be chosen early enough such that
criterion. (a) at ␶=␶1, the flow is governed by wellbore storage, (b) the
contribution ⱍ⌬p1ⱍ (Eq. 58) is sufficiently small, and (c) e␶1 is less
Starting Values. For the rate sequence y and the initial pressure or equal to the minimal elapsed time (relative to the start of its flow
p0, there are obvious choices: We initialize y to the given rate period) of any point in the pressure signal. In field examples, we
sequence q, and p0 to the maximum of the pressure sequence if all routinely use e␶1⳱10−3 hours.

December 2004 SPE Journal 381


never found rank deficiency in the matrix A at the default values of
the parameters ␯ and ␭.
We implemented two versions of our algorithm, a prototype in
Mathematica (Wolfram Inc.) and a more efficient version in C.
The C version makes use of the MESCHACH library of matrix
algebra and factorization routines.40 CPU timings for the C version
are given at the end of the field example.

Tests With Simulated Data


Figs. 1 through 7 show results for essentially the same simulated
example that was already used in Paper I. The simulated reservoir
behavior is radial flow with wellbore storage, skin, and a sealing
fault according to the model given by Agarwal et al.41 The dimen-
sionless parameters are CD⳱100, S⳱5, and a distance to the fault
of 300 wellbore radii. For the sake of generality, all other quanti-
ties in this example are dimensionless as well.
Fig. 1—Simulated example: Type curves for unbounded radial Fig. 1 shows a double logarithmic plot of the dimensionless
flow (grey) and radial flow with sealing fault (black), and inter- unit pressure drop G and the derivative dG/dlnt⳱tg(t) over time t
polated derivative type curve (black with dots at nodes). Dashed for this model (black) together with the type curves for infinite
vertical lines mark the durations of the longest flow period and radial flow (grey) and a linear interpolation of the derivative type
the entire test. curve as a thick black curve with dots at the nodes
k−1 T
3. The choice of inner nodes ␶2. . .␶n−1 is completely arbitrary ␶k = ln2 + ln , k = 1 . . . 21. . . . . . . . . . . . . . . . . . . . . . . (64)
20 2
in principle. The total number n should be large enough in order to
allow sufficient resolution of the characteristic features of the re- Fig. 3 shows simulated rates and pressure signals as functions of
sponse function, and small enough not to render the estimation time. The duration of the longest flow period is 2×104, which is
problem underdetermined. Nonuniform spacing may be desirable marked by the left vertical dashed line in Fig. 1; the duration of the
in order to render specific details of the response at higher reso- entire test is 2×105, which corresponds to the right dashed line.
lution. In practice, however, we have so far always used evenly Thus, by construction, the difference between the two reservoir
spaced nodes models emerges only at a time which is greater than the duration
of the longest individual flow period, and is therefore invisible for
k−1 conventional derivative analysis.
␶k = ␶1 + 共␶ − ␶ 兲, k = 1 . . . n, . . . . . . . . . . . . . . . . . . . . (63)
n−1 n 1 Clean data are shown in black. Four levels of measurement
errors are simulated as shown in Table 2. Here, error levels are
with n⳱20 in the simulations to be discussed later, and n⳱40 in defined as relative standard deviations:
field examples.
公m␴p 公N␴q
Implementation ␩p := , ␩q := . . . . . . . . . . . . . . . . . . . . . . . . . . . . (65)
㛳⌬p㛳2 㛳q㛳2
For the minimization of the error measure (Eq. 26) we use the
Variable Projection algorithm35,39 which is the standard algorithm An error level of 0.5% in the pressure signal would correspond to
for separable Least Squares problems. In addition to the functions the following parameter combination: Permeability 100 md, res-
F and ␷ themselves, it requires the derivatives of the residue r, (Eq. ervoir thickness 50 ft, flow rate 1,000 B/D, viscosity 0.6 cp, and
29), with respect to the nonlinear parameters x⬙⳱z. For the case of pressure uncertainty 2.5 psi. In addition to the perturbations con-
stepwise rates and linear response interpolation, expressions for sidered in Paper I, we also simulated an error level of 5% in the
ⵜzr are given in Appendix A. pressure signal.
The Variable Projection algorithm requires the solution of two Fig. 2 shows the estimated pressure derivative curves at the
linear Least Squares problems in each iteration. For these, we default parameters (␯def, ␭def)⳱(44.87, 424.8) (solid lines) and
implemented two algorithms, one based on the QR decomposition also at the parameter combination (␯GCV, ␭def) (dashed lines),
and one on Singular Value Decomposition (SVD). The advantage where ␯GCV is the GCV minimizer obtained with the results at the
of the QR decomposition is that it can be computed in situ, which default settings. The starting values z0 are shown as isolated black
makes it the method of choice for large data sets. Its only disad- dots. Except for the cases with 10% rate errors, the results are
vantage is that in case of rank deficiency it is not guaranteed to almost indistinguishable from the true type curve. This is a sub-
produce the minimum 2-norm solution. However, rank deficiency stantial improvement over earlier experiments with a regulariza-
is usually avoided due to regularization. As a check, we compute tion scheme penalizing derivatives (cf., Fig. 4 in Paper I). For data
the singular values of the matrices A and C after the last iteration. set 4, the result with default settings deviates from the true type
We encountered rank deficiency of C in a number of cases, indi- curve at late times, whereas the result with GCV-adapted weight
cating that an unregularized attempt to correct the rates based on parameter is much closer to the true type curve. The results for data
the last response iterate would be underdetermined. However, we set 5 deviate considerably from the true behavior at early and late

382 December 2004 SPE Journal


• At the default weight ␯def, and within the range investigated,
the level of regularization has hardly any effect on the rate esti-
mate. The effect on the pressure derivative estimate is much more
pronounced.
• The confidence regions around the derivative widen at early
and late times. This is to be expected in general, as these parts of
the derivative curve are sampled by fewer data points.
• Increasing the regularization parameter produces narrower
confidence regions, and thus reduces the space of possible inter-
pretation models.
• However, regularization introduces bias. This is hardly no-
ticeable at the lower levels of regularization, although at the high-
est level (Fig. 7), the pressure derivative curves are evidently too
“stiff” to fit the true reservoir behavior; likewise, the black dots
marking the mean rate estimate begin to differ from the true rate
(solid black curve). The bias in the response is especially obvious
at late times, where the curves group around the type curve for
Fig. 2—Simulated example: Type curves as in Fig. 1 with esti- infinite radial flow, thereby illustrating the dangers of using too
mated pressure derivative curves for default parameters (solid) much regularization.
and GCV adapted error weight (dashed); shading indicates er-
ror levels of input signals as per legend. Starting values shown
• The optimal level of regularization is a compromise between
as isolated black dots. these two effects; roughly speaking, the level should be chosen as
high as possible without generating visible bias. In practice, this
means that one should stop increasing the regularization once the
times and would probably not lead to a correct interpretation. Here, derivative curve looks smooth enough to be interpretable, and
a single GCV iteration does not improve the result. before its “dominant features” begin to flatten out. For the example
Fig. 4 shows the relative errors of the estimates, again with simulated here, the default level is clearly preferable to the other
default and GCV-adapted parameters. As was to be expected, the two (except for the clean data set for which the estimate always has
effect of the 5% error in the pressure signal is much more dam- zero variance; however, this is an artificial case). In general, the
aging for the pressure derivative estimate than for the rate estimate. optimal level depends on the reservoir behavior as well as on the
The effect of using the GCV-adapted weight is again most distinct error levels in the data, and finding it is a matter of trial and error.
in the derivative estimate for data set 4; the effect on the other With field data, we have often found this level to be three to four
results is relatively minor. orders of magnitude above the default given by Eq. 52.
Table 3 gives a comparison of error level estimators for the • For fixed parameters ␯ and ␭, the confidence intervals widen
data based on Eqs. 46, 50, and 51. Evidently the SVD-based pres- with increasing levels of error in the data, as is evident from Eqs.
sure error estimator comes very close to the true value, except with 42 and 43. (The difference between the two cases with 10% rate
clean data. The two rate error estimators underestimate the actual errors is only visible for small rates.) In particular, there is a
level of errors by 10 to 20%; the GCV-based estimator comes critical threshold of errors at which the choice of interpretation
closer to the truth. (A proper assessment of the rate error estimators model ceases to be unique even at the optimal level of regulariza-
would require longer rate sequences.) tion. For the example studied here, it appears that the case with 5%
A more systematic error analysis is possible on the basis of the errors in the pressure signal is clearly above this threshold.
statistical properties derived earlier. Figs. 5 through 7 show pre-
dicted expectation values and confidence regions (Eq. 44) of the
estimates obtained with the default error weight ␯def and three Field Example
different levels of regularization, namely ␭=0.01×␭def (Fig. 5), The final section presents an application of our deconvolution
␭=␭def (Fig. 6), and ␭⳱100×␭def (Fig. 7). Confidence regions are algorithm to a large set of production data from a North Sea oil
shown as shaded regions with grey levels, according to the legend well. Data were gathered only during a drill stem test of 50 hours
in the derivative plots; expectation values (which involve only the and an extended well test over the last 8,000 hours of a total of
bias, and thus depend only on the regularization level) as white 30,000 hours of production history. After the DST, the pressure
lines with black dots at the nodes in the derivative plots (respec- signal is only available again for the last 6,000 hours. This part of
tively, as black dots in the rate plots). The following observations the signal clearly indicates some unrecorded production. A build-
can be made: up test of 3,240 hours is visible at the end, which makes this an

Fig. 3—Simulated example: Rates (left) and pressure signals (right) generated with the model response shown in Fig. 1. Clean data
are in black; perturbed rate signals are in grey for 1% (dashed) and 10% error level (solid); perturbed pressure signals are shown
as dots in grey for 0.5% and in black for 5% error level.

December 2004 SPE Journal 383


Fig. 4—Simulated example: Relative errors of the rate (left) and response estimates (right) with default parameters (solid) and with
GCV adapted error weight (dashed). Legend as in Fig. 2.

ideal data set for a comparison with conventional derivative analy- As for the comparison between deconvolution and derivative
sis (Fig. 8, left plot). estimates, the shapes obtained from both types of analysis are
For deconvolution, we used the 35 rates visible in the left plot remarkably similar. To examine the reason for the vertical shift
with two different parts of the pressure signal, namely (1) the final between them at late times, we computed the multirate-type curve
buildup alone (1,293 data points after minor editing), and (2) all which conventional derivative analysis would ideally obtain if the
other buildups after 25,300 hours (9,221 data points, marked as true reservoir behavior was given by the black curve and the mea-
black dots in the right plot of Fig. 8). Grey dots mark data points sured signals were free of errors. The result is the dashed black
not contained in either set. From the drill stem test, the average curve, which is in good agreement with the conventional derivative
initial reservoir pressure is known to be p0⳱7,296 psi. The two of the last buildup (light grey dots). Hence, the shift is due to the
sets of default parameters computed from Eq. 52 are: (1) well-known distortion of multirate results in conventional deriva-
␯ d e f ⳱0.0316 (psi⭈d/bbl) 2 and ␭ d e f ⳱1.67×10 6 psi 2 ; (2) tive analysis, which is caused by the use of the superposition
␯def⳱0.0120 (psi⭈d/bbl)2 and ␭def⳱6.31×105 psi2. function.
Fig. 9 shows a comparison of deconvolution and conventional Our conclusion for this example is that deconvolution and con-
estimates. The original purpose of the final buildup test was to ventional analysis give consistent results up to a systematic bias in
establish whether the positive slope observed in the late time be- the latter. More importantly, deconvolution is able to extract the
havior of an earlier build-up (dark grey dots) was due to reservoir correct late-time behavior already from the earlier build-up data; it
boundaries. The test confirmed that this was indeed the case (light would therefore have made the final shut-in of 41⁄2 months unnec-
grey dots). These results were obtained by conventional derivative essary had it been available to the engineers.
analysis. The earlier buildup shown is in fact the longest flow Fig. 10 shows estimated confidence regions for the rate esti-
period before the final buildup for which data were available; its mate obtained from data set 2. Shading, regularization levels, and
duration is marked by the left vertical dashed line. This is the estimated error levels are as in Fig. 9 and Table 4. Despite the
maximal time interval for which conventional analysis can provide considerable size of the error margins, some of the given rates still
an estimate. The right vertical dashed line marks the time differ- lie outside these regions, which suggests that the rate errors do not
ence between the first and the last pressure sample in data set 2. follow Gaussian statistics. This view is also supported by the size
Deconvolution results obtained from the earlier buildups (data of the GCV minimizers which were of the order of 10−10 (psi⭈d/
set 2) at regularization levels of 103 and 104 times the default are bbl)2; setting the error weight to this value would amount to ig-
shown in dark grey, respectively, a lighter shade of grey; results noring the rate data.
obtained for lower levels of regularization were not smooth Fig. 11 shows the rate and pressure matches achieved with the
enough to be interpretable and are therefore not included. The deconvolution results at a regularization level of ␭⳱103␭def. The
black region marks the deconvolution result for the final build-up rate match shows that some of the rates are adapted by up to 20%;
alone (data set 1) at default regularization. the average rate match per period at this level of regularization is
Evidently the three deconvolution results are in good agree- 17.5% (see Table 4). The average pressure match per point is
ment at late times. The slight disagreement between the results approximately 200 psi, or 10% of the average RMS pressure drop
from data set 2 is entirely due to the different level of regulariza- of the buildups used for deconvolution (dark grey dots in the right
tion; naturally the result obtained with ␭/␭def⳱104 is smoother. plot). Evidently, the drawdown periods are not well matched; this
The somewhat larger deviation of the result from the final buildup is because the skin was changing with time.
alone at intermediate times may be due to a difference in wellbore The average CPU time for a single iteration with data set 2 (33
storage. rates, 9,221 pressure data points) was 13.5 seconds on a personal

384 December 2004 SPE Journal


Fig. 5—Simulated example: Expectation values and confidence intervals for pressure derivative (left) and rates (right) at the default
error weight and a regularization level of 1% of the default. Expected estimates are shown as white lines with black dots (pressure
derivative) and black dots (rate); shading of confidence intervals as per legend.

computer with an Intel Pentium III processor driven at 500 MHz certainty. At 5% pressure error and 10% rate error, there is no such
clock speed. The number of iterations required to converge (at range, and as a consequence it is possible to miss the correct
␽⳱10−5) varied from 6 at the highest to 42 at the lowest (default) interpretation of the late-time behavior.
regularization level. We have also shown an application to a large field example for
which our method yields the same late-time behavior with data
Conclusions and Discussion from different parts of the pressure signal. Moreover, the results
We have presented a new time-domain method for the deconvo- obtained from deconvolution are consistent with those obtained
lution of well-test data, which is characterized by the following from conventional analysis up to a known systematic error in the
novel features: latter.
1. A nonlinear encoding of the rate-normalized pressure deriva- In situations in which the reservoir behavior is likely to un-
tive which avoids explicit sign constraints. dergo major changes over the duration of the data set, deconvo-
2. An error measure that accounts for errors in the flow rate as lution must be used with caution. Typical examples are:
well as in the pressure signal, and thus is capable of providing a • Changing skin due to transport of solid particles into or out
joint estimate of initial pressure, rates, and pressure derivative. of the zone around the wellbore.
3. Regularization by curvature, which allows the user to control • Changing wellbore storage due to phase redistribution in the
the degree of smoothness but avoids the flattening of slopes asso- well.
ciated with regularization based on derivatives. • Liquid buildup around the wellbore during drawdowns in gas
4. Finally, to our knowledge, it is the first method to provide condensate reservoirs.
estimates for the error level of the input signals, as well as bias and • Water invasion.
confidence intervals for the results. In such circumstances, matching an entire set of production data
Our formulation includes the correction of rates and initial with a single reservoir model may neither be possible nor physi-
average pressure as a subproblem; thus, the manual correction of cally meaningful. In a recent review of our method, Levitan pro-
rates prior to analysis (though not the time synchronization of flow vided a simulated example consisting of two build-up periods with
periods with the pressure signal) is rendered unnecessary. different wellbore storage.20 Two different strategies have been
The simulated example shows that the new regularization suggested to handle this type of situation:
scheme based on curvature is a substantial improvement over ear- 1. To apply deconvolution to each flow period separately, keep-
lier schemes, which were based on derivatives and thus flattened ing initial pressure and rates fixed in each run. This is the strategy
nonzero slopes. The perturbation analysis of this example shows recommended by Levitan.
that, up to error levels of 0.5% in the pressure signal and 10% in 2. To apply deconvolution successively to increasing portions
the rate signal, there exists a range of regularization parameters for of the signal while monitoring rate and pressure match. This strat-
which the confidence regions contain the true pressure derivative egy was recently applied to data from a North Sea horizontal well
and are narrow enough to lead to a correct interpretation with affected by water invasion.28

Fig. 6—As in Fig. 5, but with regularization at the default value.

December 2004 SPE Journal 385


Fig. 7—As in Fig. 5, but with regularization at 100 times the default value.

Another problem still unresolved is the search for better (i.e., G(t) ⳱ rate-normalized pressure drop, Eq. 2, mt–1 L–4
less subjective) criteria for the selection of error weight and regu- h⳱ uniform step size between response nodes ␶k
larization parameter. Further work is needed to investigate these IN ⳱ identity matrix of size N×N
issues in greater depth and generality. k⳱ vector in curvature measure, Eq. 24
Nomenclature m⳱ number of pressure data points
mn(p) ⳱ normalized gas pseudopressure, mt–2 L–2
In order to avoid conversion factors in equations, all quantities
appearing in them are assumed to be either dimensionless or to M⳱ matrix of rate-response domain overlaps, Eq. 60
have matching units. n⳱ number of nodes for the deconvolved response
aj ⳱ start of flow period j,t N⳱ number of flow periods
A ⳱ design matrix for linearized LS estimate, Eq. 36 N(x,V) ⳱ normal distribution with mean x and covariance V
bj ⳱ end of flow period j,t p,pi ⳱ (vector of) measured pressure data, mt–2 L–2
B ⳱ design matrix for estimate with true data, Eq. C-2 p0 ⳱ average initial pressure, mt–2 L–2
c ⳱ vector in pressure variance estimate, Eq. 46 p* ⳱ true pressure sequence, mt–2 L–2
c0, c1 ⳱ coefficients in initial reservoir response, Eq. 54 q,qj ⳱ (vector of) measured rate data, L3/t
C(z) ⳱ design matrix for pressure match, Eq. 16 Q(t) ⳱ rate signal as a function of time, L3/t
C* ⳱ matrix C(z) for the true response z*, Eq. 35 r(x⬘,x⬙) ⳱ residue of the separable LS problem, Eq. 29
Cij(z) ⳱ components of the matrix C(z), Eq. 16 R⳱ confidence region of estimate x̂ , Eq. 44
Cijk(z) ⳱ auxiliary function for evaluation of C(z), Eq. A-3 t⳱ time
d ⳱ number of rows in matrix D ti ⳱ time at which pressure sample pi was measured
D ⳱ matrix in curvature measure, Eq. 24 T⳱ test duration, t
DC ⳱ diagonal matrix in SVD of C* u⳱ vector (1,1) ∈⺢2
*
Dijk(z) ⳱ auxiliary function for evaluation of the Jacobian of U, V ⳱ orthogonal matrices in SVD of C*, after Eq. 46
C(z), Eq. A-7 w(t) ⳱ weight function (in Gajdica et al.’s formulation10)
E ⳱ error measures for estimate of initial pressure, rates, x, xi ⳱ (vector of) parameters estimated in the joint
and response linearized LS problem,
E0 ⳱ initial value of the error measure x ⳱ (p0,y,z), Eq. 36
E1 ⳱ error measure for estimate of response only x⬘ ⳱ (p0,y): parameters in which the residue r is linear
E2 ⳱ error measure for estimate of initial pressure and x⬙ ⳱ z: parameters in which the residue r is nonlinear
rates only x* ⳱ vector of true initial pressure, rates, and response
F(x⬙) ⳱ design matrix in separable nonlinear LS residue coefficients
g (t) ⳱ reservoir impulse response, Eq. 2, mt–2 L–4 y,yj ⳱ (vector of) estimated rates, L3/t

Fig. 8—Field example: Entire data set (except for an initial 50 h test, left) and close-up of earlier build-ups used for deconvolution
(right, enlarged). Grey dots mark unused pressure data.

386 December 2004 SPE Journal


Fig. 9—Pressure derivative estimates obtained from conven- Fig. 10—Confidence regions for the rates estimated from data
tional analysis and from deconvolution. set 2. Shading as in Fig. 9 and Table 4.

y* ⳱ vector of true rates, L3/t ⌼m ⳱ vector of size m with each component equal to 1
z,zk ⳱ (vector of) response coefficients Z(␶k) ␸k ⳱ angle between line segments joined at node ␶k (see
z* ⳱ true values for vector z Fig. 12)
Z(␶) ⳱ reservoir response for the diagnostic plot, Eq. 11 ␹2(p0,y)⳱ negative logarithm of the likelihood function for the
Z0(␶) ⳱ initialization for reservoir response Z(␶), Eq. 54 estimate of initial pressure and rates with fixed
␣ ⳱ “energy” threshold (in Baygün’s formulation15) response, Eq. 49
␣k, ␤k ⳱ coefficients in linear response interpolation Eq. 18 ␺k(␶) ⳱ response interpolants, Eq. 13
␥(t) ⳱ logarithmic pressure derivative, Eq. 4 u* ␷ ⳱ convolution product of functions u and ␷, Eq. 1
⌫ ⳱ Jacobian of z哫C(z)y, at z=z*, Eq. 35 (u, ␷) ⳱ scalar product of functions u and ␷, Eq. 8
␦,␦j ⳱ (vector of) absolute rate errors, L3/t 㛳u㛳n ⳱ Ln-norm of a function u, Eq. 5
␦ij ⳱ Kronecker symbol: 1 if i=j, zero otherwise ū ⳱ spectral transform of function u
⌬(t) ⳱ rate error signal as function of time, L3/t, Eq. 21 ũ ⳱ spectral transform of function u truncated to interval
⌬p(t) ⳱ pressure drop signal, mt–2 L–2 [0, T], Eq. 10
⌬p,⌬pi ⳱ (vector of) pressure drop data, mt–2 L–2 㛳␷㛳2 ⳱ ᐉ2-norm of a vector ␷, Eq. 22
␧,␧i ⳱ (vector of) absolute pressure data errors, mt–2 L–2 diag(␷) ⳱ diagonal matrix with diagonal entries ␷1, ␷2, …
␩p, ␩q ⳱ error levels of pressure and rate signals, Eq. 65 ␷ˆ ⳱ estimate of ␷
␪j(t) ⳱ rate interpolants, Eq. 13 ⺕{␷} ⳱ expectation value of ␷
␽ ⳱ threshold in stopping criterion, Eq. 62 ⺒{␷} ⳱ ⺕ {(␷−␷ˆ } bias of estimate ␷ˆ
␬(z) ⳱ curvature measure, Eq. 24 and Appendix B cov{␷} ⳱ ⺕ {(␷−␷ˆ )(␷−␷ˆ )T} covariance matrix of vector ␷
␭ ⳱ regularization parameter, Eq. 23 ␷T ⳱ transpose of vector ␷
␭def ⳱ default value for regularization parameter, Eq. 52 AT ⳱ transpose of matrix A
␮ ⳱ midpoint of overlap interval, Appendix A A† ⳱ Moore-Penrose inverse35 of matrix A
␮j(t) ⳱ rate-response overlap measure, Eq. 57
␯ ⳱ relative error weight, Eq. 23 Acknowledgments
␯def ⳱ natural unit for relative error weight, Eq. 52ⱍ This work was supported by BP plc, Schlumberger, Norsk Hydro,
␳ ⳱ radius of overlap interval, Appendix A and Conoco not only financially, but also by giving us their feed-
␳(␶) ⳱ response autocorrelation function (in Baygün’s back. We are particularly grateful to Michael Levitan of BP plc
formulation15), Eq. 7 and Professor Gene Golub (Dept. of Computer Science, Stanford
␴,␴i ⳱ (vector of) standard deviations of estimates for x,xi U.) for stimulating discussions.
␴2p ⳱ variance of the pressure error, m2t–4L–4
References
␴ˆ 2P ⳱ estimated variance of the pressure error, m2t–4L–4
1. Hutchinson, T.S. and Sikora, V.J.: “A Generalized Water-Drive Analy-
␴2q ⳱ variance of the rate error, L6t–2
sis,” Trans., AIME (1959) 216, 169.
␴ˆ 2q ⳱ estimated variance of the rate error, L6t–2
2. Katz, D.L., Tek, M.R., and Jones, S.C.: “A Generalized Model for
␷k ⳱ vertex of the response graph at ␶k, see Fig. 12 Predicting the Performance of Gas Reservoirs Subject to Water Drive,”
␷(x⬙) ⳱ vector in separable nonlinear LS residue, Eq. 26 paper SPE 428 presented at the 1962 SPE Annual Meeting, Los An-
␶ ⳱ natural logarithm of time geles, 7–10 October.
␶k ⳱ nodes for logarithmic response interpolation 3. Coats, K.H. et al.: “Determination of Aquifer Influence Functions
From Field Data,” Trans., AIME (1964) 231, 1417.
4. Jargon, J.R. and van Poollen, H.K.: “Unit Response Function From
Varying-Rate Data,” JPT (August 1965) 965; Trans., AIME, 234.
5. Bostic, J.N., Agarwal, R.G., and Carter, R.D.: “Combined Analysis of
Postfracturing Performance and Pressure Buildup Data for Evaluating
an MHF Gas Well,” JPT (October 1980) 1711.
6. Kučuk, F.J. and Ayestaran, L.: “Analysis of Simultaneously Measured
Pressure and Sandface Flow Rate in Transient Well Testing,” JPT
(February 1985) 323.
7. Thompson, L.G. and Reynolds, A.C.: “Analysis of Variable-Rate Well-
Test Pressure Data Using Duhamel’s Principle,” SPEFE (October
1986) 453.

December 2004 SPE Journal 387


Fig. 11—Rate match (left) and pressure match (right). Given rates are in black, deconvolved rates in grey; pressure data used for
deconvolution are in dark grey, remaining pressure data in light grey, pressure signal from deconvolved response and fitted rates
in black, and the initial average pressure is shown as black dot.

8. Thompson, L.C., Jones, J.R., and Reynolds, A.C.: “Analysis of Pres- 19. von Schroeter, T., Hollaender, F., and Gringarten, A.C.: “Analysis of
sure Buildup Data Influenced by Wellbore Phase Redistribution,” Well Test Data From Permanent Downhole Gauges by Deconvolu-
SPEFE (October 1986) 435. tion,” paper SPE 77688 presented at the 2002 SPE Annual Technical
9. Kuchuk, F.J.: “Applications of Convolution and Deconvolution to Conference and Exhibition, San Antonio, Texas, 29 September–2 Oc-
Transient Well Tests,” SPEFE (December 1990) 375; Trans., AIME, tober.
289. 20. Levitan, M.M.: “Practical Application of Pressure-Rate Deconvolution
to Analysis of Real Well Tests,” paper SPE 84290 presented at the
10. Gajdica, R.J., Wattenbarger, R.A., and Startzman, R.A.: “A New
2003 SPE Annual Technical Conference and Exhibition, Denver, 5–8
Method of Matching Aquifer Performance and Determining Original
October.
Gas in Place,” SPERE (August 1988) 985.
21. van Everdingen, A.F. and Hurst, W.: “The Application of the Laplace
11. Roumboutsos, A. and Stewart, G.: “A Direct Deconvolution or Con-
Transformation to Flow Problems in Reservoirs,” Trans., AIME (1949)
volution Algorithm for Well Test Analysis,” paper SPE 18157 pre-
186, 305.
sented at the 1988 SPE Annual Technical Conference and Exhibition,
22. Al-Hussainy, R., Ramey Jr., H.J., and Crawford, P.B.: “The Flow of
Houston, 2–5 October.
Real Gas Through Porous Media,” JPT (May 1966) 624.
12. Mendes, L.C.C., Tygel, M., and Corrèa, A.C.F.: “A Deconvolution 23. Bøe, A., Skjaeveland, S.M., and Whitson, C.H.: “Two-Phase Pressure
Algorithm for Analysis of Variable Rate Well Test Pressure Data,” Test Analysis,” SPEFE (December 1989) 604; Trans., AIME, 287.
paper 19815 presented at the 1989 SPE Annual Technical Conference
24. Bourdet, D. et al.: “A New Set of Type Curves Simplifies Well Test
and Exhibition, San Antonio, Texas, 8–11 October.
Analysis,” World Oil (1983) 196, 95.
13. Kuchuk, F.J., Carter, R.G., and Ayestaran, L.: “Deconvolution of Well- 25. Odeh, A.S. and Jones, L.G.: “Pressure Drawdown Analysis, Variable-
bore Pressure and Flow Rate,” SPEFE (March 1990), 53. Rate Case,” JPT (August 1965) 960.
14. Bourgeois, M.J. and Horne, R.N.: “Well-Test-Model Recognition with 26. Bourdet, D., Ayoub, J.A., and Pirard, Y.M.: “Use of Pressure Deriva-
Laplace Space,” SPEFE (March 1993) 17. tive in Well-Test Interpretation,” SPEFE (June 1989) 293; Trans.,
15. Baygün, B., Kuchuk, F.J., and Arikan, O.: “Deconvolution Under Nor- AIME, 287.
malized Autocorrelation Constraints,” SPEJ (September 1997) 246. 27. Daungkaew, S., Hollaender, F., and Gringarten, A.C.: “Frequently
16. Onur, M. and Reynolds, A.C.: “Numerical Laplace Transformation of Asked Questions in Well Test Analysis,” paper SPE 63077 presented at
Sampled Data for Well-Test Analysis,” SPEREE (June 1998) 268. the 2000 SPE Annual Technical Conference and Exhibition, Dallas,
17. Gilly, P. and Horne, R.N.: “Analysis of Pressure/Flow-Rate Data With 1–4 October.
the Pressure-History-Recovery Method,” SPEREE (August 1999) 314. 28. Gringarten, A.C. et al.: “Use of Downhole Permanent Pressure Gauge
18. von Schroeter, T., Hollaender, F., and Gringarten, A.C.: “Deconvolu- Data to Diagnose Production Problems in a North Sea HorizontalWell,”
tion of Well Test Data as a Nonlinear Total Least Squares Problem,” paper 84470 presented at the 2003 SPE Annual Technical Conference
paper SPE 71574 presented at the 2001 SPE Annual Technical Con- and Exhibition, Denver, 5–8 October.
ference and Exhibition, New Orleans, Louisiana, 30 September–3 Oc- 29. Wing, M.G.: A Primer on Integral Equations of the First Kind, Society
tober. for Industrial and Applied Mathematics (SIAM), Philadelphia (1991).
30. Golub, G.H., Hansen, P.C., and O’Leary, D.P.: “Tikhonov regulariza-
tion and Total Least Squares,” SIAM J. Matrix Anal. Appl. (1999) 21,
185.
31. Percival, D.B. and Walden, A.T.: Spectral Analysis for Physical Ap-
plications, Cambridge U. Press, Cambridge, U.K. (1993).
32. Stehfest, H.: “Algorithm 368: Numerical inversion of Laplace trans-
forms,” Comm. Assoc. Computing Machinery (1970) 13, 47.
33. Dong, C.W.: “A regularization method for the numerical inversion of
the Laplace transform,” SIAM J. Num. Anal. (1993) 30, 759.
34. Golub, G.H. and van Loan, C.F.: Matrix Computations, third edition,
The Johns Hopkins U. Press, Baltimore (1996).
35. Björck, Å.: Numerical Methods for Least Squares Problems, Society for
Industrial and Applied Mathematics (SIAM), Philadelphia (1996).
36. Press, W.H. et al.: Numerical Recipes in PASCAL. The Art of Scientific
Computing, Cambridge U. Press, Cambridge, U.K. (1989).
37. Golub, G.H., Heath, M., and Wahba, G.: “Generalized Cross-
Validation as a Method for Choosing a Good Ridge Parameter,” Tech-
Fig. 12—Angle between adjoining line segments. nometrics (1979) 21, 215.

388 December 2004 SPE Journal


38. Hansen, P.C. and O’Leary, D.P.: “The use of the L-curve in the regu- It follows from Eq. A-4 that


larization of discrete ill-posed problems,” SIAM J. Scientific Comput-
⭸Z ␦ if ␶ ⱕ ␶
ing (1993) 15, 1487. 共␶兲 = ␺k,1共␶兲 if ␶ ⱕ ␶1 ⱕ ␶ . . . . . . . . . . . . . . . . . . . . . . . . (A-6)
39. Golub, G.H. and Pereyra, V.: “The differentiation of pseudoinverses ⭸zk k 1 n

and nonlinear least squares problems whose variables separate,” SIAM and therefore from Eq. A-1 that
J. Num. Anal. (1973) 10, 413.
␶1 lnT
⭸Cij a1
兰 兰␪ 共t − e 兲␺ 共␶兲e
40. Stewart, D.E. and Leyk, Z.: “MESCHACH, version 1.2b,” (1994)
http://www.netlib.org. = e ␦k,1 ␪j共ti − e␶兲e␶␤kd␶ + ␶ ␶␤k
d␶.
⭸zk j i k
41. Agarwal, R.G., Al-Hussainy, R., and Ramey Jr., H.J.: “An Investiga- −⬁ ␶1
tion of Wellbore Storage and Skin Effect in Unsteady Liquid Flow. I: Here, the first term is zero unless k⳱1, in which case the first
Analytical Treatment,” SPEJ (September 1970) 279; Trans., AIME, integral is Cij1(z). In the evaluation of the second integral, the cases
249. k⳱1 and k⳱n must be treated separately, as in these cases only
half of the support of ␺k overlaps with the integration domain.
Appendix A—Formulae for Stepwise Rates and
Defining
Linear Response Interpolation
␶k
This appendix gives expressions for the matrix C (Eq. 16), and its
derivatives for the case in which the rates are modeled as stepwise Dijk共z兲 = 兰 ␶␪ 共t − e 兲e
j i
␶ ␶␤k
d␶ = 兰␶e ␶␤k
d␶, . . . . . . . . . . . . . . . (A-7)
constant, and the interpolation scheme for the response is linear in ␶k−1 Iijk
the response coefficients zk. If the rate interpolants are given by one obtains
Eq. 14, then contributions to the component
⭸Cij ␣1 Dij,2共z兲 − ␶2Cij,2共z兲
lnT = e Cij1共z兲 + e␣2 . . . . . . . . . . . . . . . . (A-8)
Cij共z兲 = 兰␪ 共t − e 兲e
−⬁
j i
␶ Z共␶兲
d␶ . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-1)
⭸z1
for k⳱1,
␶1 − ␶2

arise only from the subset of the domain of integration where ⭸Cij ␣k Dijk共z兲 − ␶k−1Cijk共z兲 ␣k+1 Dij,k+1共z兲 − ␶k+1Cij,k+1共z兲
␪j(ti−e␶)⳱1. We split this subset into a sequence of nonoverlap- =e +e
⭸zk ␶k − ␶k−1 ␶k − ␶k+1
ping (possibly empty) intervals Iijk:⳱{␶∈(␶k−1,␶k):␪j(ti−e␶)⳱1}, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-9)
k⳱1. . .n. Then, it follows from Eqs. A-1, 17, and 18 that
for k⳱2…n−1, and
n

Cij共z兲 = 兺e
k=1
␣k
Cijk共z兲, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-2) ⭸Cij ␣n Dijn共z兲 − ␶n−1Cijn共z兲
⭸zn
=e
␶n − ␶n−1
. . . . . . . . . . . . . . . . . . . . . . . . (A-10)

where for k⳱n. If cases are distinguished as before, the integral in Eq.
␶k A-7 evaluates to

兰 ␪ 共t − e 兲e ␶ ␶␤k
兰e ␶␤k if Iijk = ⭋


Cijk共z兲 = j i d␶ = d␶. . . . . . . . . . . . . . . . . . (A-3) 0
␶k−1 Iijk Dijk共z兲 = 共b Ⲑ ␤k − 1 Ⲑ ␤k2兲eb␤k if a = −⬁
Now each of the sets Iijk is either empty, a semi-infinite interval 2␮␳ if a ⫽ −⬁,␤k = 0
(–⬁,b], or a finite interval of the form [a,b]⳱[␮−␳,␮+␳] with
midpoint ␮ and radius ␳. Accordingly, one obtains and finally to
2 ␮␤k
if Iijk = Ø


0 Dijk共z兲 = e 兵共␮ − ␤k−1兲sinh共␳␤k兲 + ␳ cosh共␳␤k兲其,
␤k−1eb␤k if a = −⬁ ␤k
Cijk共z兲 = 2␳, if a ⫽ −⬁ ␤k = 0 if a⫽–⬁ and ␤k⫽0.
2␤kⳮ1 e␮␤k sinh共␳␤k兲 if a ⫽ −⬁, ␤k ⫽ 0.
Appendix B—Curvature Measure
Note that a⳱–⬁ can only occur for k⳱1 (i.e., before the first
response node, where we assume unit slope, ␤1⳱1). For the sake of generality, we shall consider an arbitrary spacing
The Variable Projection algorithm also requires the Jacobian of of nodes. For a piecewise linear function, the curvature of its graph
the residue r(x⬘,x⬙) with respect to the nonlinear parameters x⬙=z. is localized at its nodes; an obvious measure of the overall curva-
Its components are ture is

⭸ri
⭸zk
= 再 − 兺 N
j=1yj⭸Cij ⭸zk
0
Ⲑ if 1 ⱕ i ⱕ m
if m ⬍ i ⱕ m + N
公␭Di−m−N,k if i ⬎ m + N.
␬= 冑兺 n−1

i=1
␸i2, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-1)

where ␸i is the angle between the segments joined at node ␶i. This
To compute the derivatives ⭸Cij /⭸zk, it is convenient to write the angle is given by
response interpolation in the form*
||共␷i+1 − ␷i兲 ∧ 共␷i − ␷i−1兲||2

Z共␶兲 = 再兺 n
z1 + ␶ − ␶1 if ␶ ⱕ ␶1
k=1zk␺k共␶ 兲 if ␶1 ⱕ ␶ ⱕ ␶n . . . . . . . . . . . . . . . . . (A-4)
sin ␸i =
||␷i+1 − ␷i||2||␷i − ␷i−1||2
, . . . . . . . . . . . . . . . . . . . . . . (B-2)

where ˆ denotes the vector product (or cross product) in 3D, and
where the ␷i are the vertices of the graph of z. The situation is schemati-
cally shown in Fig. 12. We model wellbore storage by assuming
␶ − ␶k−1


unit slope before the first node:
if ␶k−1 ⱕ ␶ ⱕ ␶k
␶k − ␶k−1 ||共␷2 − ␷1兲 ∧ u||2
␺k共␶兲 = ␶ − ␶k+1 . . . . . . . . . . . . . . . . . . . (A-5) sin ␸1 = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-3)
if ␶k ⱕ ␶ ⱕ ␶k+1 ||␷2 − ␷1||2||u||2
␶k − ␶k+1
where u⳱(1,1) is tangent to the first (infinite) segment. To bring
0 if otherwise the curvature measure into a form suitable for regularization pur-
poses, we approximate ␸i in Eq. B-1 by its sine, set the z compo-
nents of the vectors appearing in the denominators in Eqs. B-2 and
* The slight difference between Eqs. A-4 and 13 is due to our assumption of a fixed slope B-3 to 0, and replace 㛳u㛳2 in Eq. B-3 by 1. The combined effect of
before the first node; this assumption was not made in Eq. 13. these approximations is that we can write

December 2004 SPE Journal 389


␬ = ||Dz − k||2, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-4)

冤 冥
0 0 0
where J := A A − B B =
T T 0 vIN 0 , . . . . . . . . . . . . . . . . . . . (C-3)
0 0 ␭DTD
k = 共1,0, . . . ,0兲 ∈⺢ n−1
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-5)
one obtains that*
and D is a matrix of size (n−1)×n with first row given by
ATA共x̂ − x*兲 = ATb − BTc − Jx*


−共␶2 − ␶1兲 − 1 if j = 1,
D1j = 共␶2 − ␶1兲 − 1 if j = 2, . . . . . . . . . . . . . . . . . . . (B-6a)

冤 冥
ϒmT␧
0 if j = 3 . . . n, −C*T␧ − v␦
= .
and rows i⳱2…n−1 given by
−⌫ ␧ + ␭D 共k − Dz*兲
T T


共␶i − ␶i−1兲−1 if j = i − 1,
Eq. 40 now follows by taking expectations and using the assump-
␶i+1 − ␶i−1 tions that E{␧}⳱E{␦}⳱0 and that ATA is invertible.
if j = i,
Dij = 共␶i+1 − ␶i兲共␶i − ␶i−1兲 . . . . . . . . . . . . . . . (B-6b)
共␶i+1 − ␶i兲 − 1 if j = i + 1,
SI Metric Conversion Factors
0 otherwise
1 bbl ⳱ 1.589 873 × 10−1 m3
These formulae are valid for any spacing of the nodes ␶i. For 1 psi ⳱ 6.89 4757 × 103Pa
uniform spacing with step size h, rows 2 to n−1 of D reduce to the
well-known discrete approximation of a second derivative operator Thomas von Schroeter is a Research Associate at the Dept. of
with –2h–1 in the diagonal and h–1 in the sub- and superdiagonal. Earth Science and Engineering, Imperial College, London.
This curvature measure is asymptotically invariant under sub- e-mail: t.schroeter@imperial.ac.uk. His research interests in-
division of node intervals in the sense that, for a sufficiently in- clude fluid mechanics, numerical analysis, and signal process-
terpolated curve, additional nodes would not pick up additional ing. He holds a Diplom in Physics from the U. of Göttingen,
curvature. Thus, contrary to a statement made in Paper II, dividing Germany, and a D. Phil. Degree from the U. of Oxford. Florian
D and k by the Frobenius norm of D, viz. Hollaender is working for Schlumberger in the position of
Schlumberger Chair Professor at the Dept. of Chemical and
||D||F := 冑兺 i,j
Dij2, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-7)
Petroleum Engineering of the U. of the United Arab Emirates at
Al Ain. e-mail: hollaender@abu-dhabi.oilfield.slb.com. He
graduated from the École Nationale Superieure des Mines de
breaks the invariance. (Proof: For uniform spacing of n nodes over Nancy and holds a PhD degree in Petroleum Engineering from
a constant interval, h∼n–1 and 㛳D㛳F2∼6nh−2∼n3, so 㛳D㛳F∼n3/2.) Ac- Imperial College, London. Alain C. Gringarten holds the Chair
cordingly, the required value of ␭ for a given data set would of Petroleum Engineering at Imperial College, London, where
depend on n. However, as we keep the number and spacing of he is also director of the Centre for Petroleum Studies. e-mail:
a.gringarten@imperial.ac.uk. Prior to joining Imperial in 1997,
nodes fixed in each example, this dependence is irrelevant for our he held a variety of senior technical and management posi-
purposes, and we shall keep the scaling for the sake of consistency tions with Scientific Software-Intercomp; Schlumberger; and
with Paper II. Thus, we report the value of ␭ for the curvature the French Geological Survey in Orléans, France. His research
measure ␬/㛳D㛳F instead of ␬. interests include fissured fluid-bearing formations, fractured
wells, gas condensate and volatile oil reservoirs, high and low-
Appendix C—Bias of the Linearized Parameter enthalpy geothermal energy, hot dry rocks, and radioactive
Estimate waste disposal. He has published over 60 technical papers and
To derive Eq. 40, we first remark that the true parameter vector was responsible for many advances in well-test interpretation.
He holds MSc and PhD degrees in petroleum engineering from
x*⳱(p0,y*,z*) is an exact solution of the deconvolution problem
Stanford U. and an engineering degree from the École Cen-
for the true pressure sequence p*⳱p–␧ and the true rate sequence trale Paris, France. Gringarten is a recognized expert in well-
y* and hence minimizes the rate match test analysis and was the recipient of the SPE John Franklin Carll
Award for 2003 and the SPE Formation Evaluation Award for
E*共␣,y,z兲 = ||␣ϒm − p* − C*y − ⌫共z − z*兲||22. . . . . . . . . . . . . (C-1) 2001. A member of SPE since 1969, he was elected a Distin-
guished Member in 2002.
[In fact, E*(p0,y*,z*)⳱0.] Thus, both estimate and true solution
satisfy normal equations,
ATAx̂ = ATb, BTBx* = BTc, . . . . . . . . . . . . . . . . . . . . . . . . . . . . (C-2)
where B⳱[ϒm,−C* ,−⌫] and c=p*−⌫z*. Subtracting the two nor- * In the conference version,19 the term –⌫T␧ was missing in the following equation. How-
mal equations from each other and putting ever, as this term has zero expectation, the result for the bias is unchanged.

390 December 2004 SPE Journal

You might also like