Anda di halaman 1dari 16

SPE 77691

Effective Well and Reservoir Evaluation Without the Need for Well Pressure History
B. D. Poe Jr., SPE, Schlumberger
Copyright 2002, Society of Petroleum Engineers Inc.
This paper was prepared for presentation at the SPE Annual Technical Conference and
Exhibition held in San Antonio, Texas, 29 September2 October 2002.
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, as
presented, have not been reviewed by the Society of Petroleum Engineers and are subject to
correction by the author(s). The material, as presented, does not necessarily reflect any
position of the Society of Petroleum Engineers, its officers, or members. Papers presented at
SPE meetings are subject to publication review by Editorial Committees of the Society of
Petroleum Engineers. Electronic reproduction, distribution, or storage of any part of this paper
for commercial purposes 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 where and by whom the paper was presented. Write Librarian, SPE, P.O.
Box 833836, Richardson, TX 75083-3836, U.S.A., fax 01-972-952-9435.

Abstract
This paper presents a practical and useful analysis technique
for effectively accomplishing a rigorous superposition-in-time
(convolution) analysis of the transient drawdown performance
of oil and gas wells that does not require that the flowing
pressures (wellhead or bottom hole) be known for every flow
rate data point in the production history. If at least one
flowing pressure point is available at any point in time during
the production history of the well, a unique analysis of the
production data for the evaluation of the reservoir effective
permeability, well drainage area, and well parameters such as
unfractured well radial flow steady-state skin effect, fractured
well effective fracture half-length and average fracture
conductivity, or horizontal well effective length in the pay
zone may be obtained. Even if no flowing pressures are
available at all during the production history of the well,
estimates of these well and reservoir parameters can still be
obtained with reasonable accuracy using the analysis
techniques presented in this paper.
The general evaluation procedures presented in this paper
are applicable for all well types, including flowing wells and
wells with artificial lift systems (i.e., conventional beam
pumping units and electrical submersible pumps (ESP)). The
theoretical basis for a rate-transient analysis of the production
performance of a vertically fractured well is presented in detail
in this paper. Application of the proposed general analysis
procedure to other well types (unfractured vertical/slanted
wells, horizontal wells, etc.) may also be found in a similar
manner as presented in this paper. The theoretical basis for a
pressure-transient based solution for a vertically fractured well
is also available in the references.
In the relatively short period of time that this new analysis
technique has been available, it has radically changed the way

that engineers may now rigorously analyze the production


performance of oil and gas wells to evaluate the well and
reservoir properties from the wells production performance,
as well as greatly expanding the number and types of wells
that can now be considered. The new analysis techniques
presented in this paper therefore represent a step-change in the
petroleum industrys technical capabilities for the analysis of
the production performance of oil and gas wells over that
which existed prior to this work.
Introduction
One of the most common problems encountered by a
production or reservoir engineer when analyzing an oil or gas
wells production history to evaluate the formation and
completion system properties (by inversion of the production
history record of the well), is the lack of a complete data
record with which to employ a conventional convolution
analysis. The variable that is most commonly not recorded
that is required for a conventional convolution analysis of the
production history of the well is the well flowing pressure. An
analysis technique is presented in the following sections of
this paper that will remove this limitation in a convolution
analysis of the production data record of a vertically fractured
well. Similar analysis techniques have also been developed
for other types of wells1.
The flow rates of the hydrocarbon phases (oil and gas) of a
well are generally known with a reasonably high degree of
accuracy, since the oil and gas production directly correspond
to revenue for the operator and royalty owners. Custody
transfers also dictate that the production of these fluids be
monitored as accurately as possible. The exception to this
general statement is the case where the production of several
wells are processed at a single separation facility and the
actual production of any single well in the system may be
estimated from the combined system production with periodic
tests of the individual wells in the system.
Often known with a much lower degree of certainty is the
water production from the well, since this variable generally
represents an operating cost of the well, it is commonly not
recorded or reported with a great deal of precision. Often the
most reliable way of ascertaining a representative water
production history for a producing well is through the use of
run tickets for the loads of water transported to disposal wells.
The effect of a varying flow rate and sandface flowing
pressure history of a well on its dimensionless wellbore

B. POE

pressure solution at a point in time of interest has been


established with the Faltung Theorem as employed by van
Everdingen and Hurst2. The general form of the well-known
convolution relationship that accounts for the superpositionin-time effects of a varying sandface pressure and flow rate on
the dimensionless wellbore pressure transient behavior of a
well is given by Eq. 1.
tD

pwD ( tD ) = qD ( ) pD ' ( tD ) d

(1)

The pressure transient behavior of a well with a


varying flow rate and flowing pressure history can be readily
evaluated using Eq. 1 for specified terminal flow rate
(Neumann) inner boundary condition transients such as
constant flow rate drawdown or injection transients, or shut-in
well sequences such as pressure buildup or falloff transients.
The development of a pressure-transient based analysis that is
analogous to the rate-transient analyses presented in the
remainder of this paper may be found in Ref. 1.
As fully discussed in Ref. 3, the most appropriate inner
boundary condition solution to use for the analysis of the
production history of a well corresponds to that of a specified
terminal pressure (Dirichlet) inner boundary condition.
Therefore, further discussion of the pressure-transient solution
based analyses will be abbreviated at this point and an
example of its application for the analysis of a vertically
fractured well is provided in Ref. 1 for the sake of
completeness only. It has been included in this discussion
simply due to its historical significance and to demonstrate the
similarity of the form of the two relationships. More
importantly, one thing that the practicing engineer should be
cognizant of is the fact that the dimensionless flow rate and
pressure appearing in Eq. 1 do not equate to the dimensionless
flow rate and pressure of the rate-transient based analyses
presented in the balance of this paper.
Equation 2 gives the corresponding dimensionless ratetransient behavior of a well with a varying flow rate and
sandface flowing pressure history3.
tD
0

tD

qwD ( tD ) = pD ( )qD ( tD ) d

i =1
n >1

+ qD ( tDn tDn 1)

(4)

The corresponding rate-transient solution dimensionless


cumulative production of a well with a varying flow rate and
sandface pressure production history can also be evaluated
using a discrete time approximation3 with Eq. 5.

QpD ( tDn ) = pDi QpD ( tDn tDi 1) QpD ( tDn tDi )


i =1
n >1

+ QpD ( tDn tDn 1)

(3)

Using either the pressure-transient (Eq. 1) or rate-transient


(Eq. 3) convolution integrals to account for the varying flow
rate and sandface pressure history of a well, a discrete time
approximation of the convolution integral can be derived to
permit the analysis of a varying flow rate and sandface
pressure production history. The corresponding rate-transient
convolution integral approximation of the dimensionless well
flow rate is given in Eq. 4.

(5)

The dimensionless pressure appearing in the superpositionin-time relationships of Eqs. 4 and 5 is defined for oil and gas
reservoirs with Eqs. 6 and 7, respectively.

pDi =
pDi =

pi pwf ( ti )
pi pwf ( tn )

(6)

pp ( pi ) pp ( pwf ( ti ) )

(7)

pp ( pi ) pp ( pwf ( tn ) )

The well bore dimensionless flow rate is defined for oil


and gas reservoirs in conventional oilfield units with Eqs. 8
and 9, respectively.

qwD =
qwD =

141.205 qo ( t ) o Bo
ko h ( pi pwf )

(8)

50299.5 psc T qg ( t )

kg h Tsc ( pp ( pi ) pp ( pwf ) )

(9)

The dimensionless cumulative production of oil and gas


reservoirs is defined in conventional oilfield units as given in
Eqs. 10 and 11, respectively.

QpD ( tn ) =

(2)

With a substitution of variables, the rate-transient


convolution integral can be written in the more readily
amenable form presented in Eq. 3.
'

n 1

qwD ( tDn ) = pDi qD ( tDn tDi 1) qD ( tDn tDi )

n 1

qwD ( tD ) = qD ( tD ) pD ' ( ) d

SPE 77691

QpD ( tn ) =

Np ( tn ) Bo

1.11909 ct h Lc 2 ( pi pwf ( tn ) )
318313 psc T Gp ( tn )

(10)

h gct ( tn ) Tsc Lc 2 pp ( pi ) pp ( pwf ( tn ) )

(11)
The dimensionless time that corresponds to a given value

of dimensional time ( tn ) for oil and gas reservoir analyses is


defined with Eqs. 12 and 13, respectively.

tD ( tn ) =

0.000263679 ko tn
o ct Lc 2

(12)

tD ( tn ) =

0.000263679 kg ta ( tn )
Lc 2

(13)

SPE 77691

EFFECTIVE WELL AND RESERVOIR EVALUATION WITHOUT THE NEED FOR WELL PRESSURE HISTORY

The system characteristic length

( Lc )

appearing in Eqs.

10 through 13 is dependent upon the system under


consideration. In an unfractured vertical well analysis it is
equal to the wellbore radius (half the wellbore diameter, not
necessarily or generally equal to the hole size). An apparent
(or effective) well bore radius is also commonly used as the
system characteristic length in unfractured vertical well
decline analyses, particularly in cases where the well has been
stimulated to improve its productivity, resulting in a negative
steady state skin effect. In this case, the apparent well bore
radius (or the system characteristic length) is simply the well
bore radius multiplied by the exponential function value of the
negative of the steady state skin effect.
In a vertically fractured well analysis, the system
characteristic length is the fracture half-length, or half of the
total effective fracture length in the system. Similarly, in a
horizontal well analysis the system characteristic length is
equal to half of the total effective wellbore length in the
pay zone.
The evaluation of the pseudotime integral transformation is
addressed in detail in Ref. 4 and a detailed discussion of this
topic need not be given in this paper. However, suffice it to
say that significant care must be employed in lowpermeability gas reservoir analyses so that this integral
transformation is accurately and properly evaluated.
With
these
rate-transient
analysis
fundamental
relationships established, we turn our attention next to the
development of a practical means of estimating the
superposition-in-time function values of production history
data points for which the flowing sandface (or wellhead)
pressure are not available. For a production history data point
that has the flowing wellhead pressure and well flow rates
recorded, the corresponding bottom hole wellbore and
sandface flowing pressures can be readily be estimated using
the industry-accepted wellbore pressure traverse and
completion pressure loss models documented in Ref. 5.
When the wellhead flowing pressure is not available at a
production data point, and bottom hole pressure measurements
are also not available, a conventional convolution analysis of
the type prescribed by Eqs. 4 and 5 is not possible without
guessing (or in some way roughly estimating) what the
missing sandface flowing pressure should have been at that
point in time in the production history.
Palacio and Blasingame6, using the material balance time
function that was derived and first used by McCray7, proposed
an alternative solution to this problem. The material
balance equivalent time function casually appears similar to
the Horner approximation commonly used for the evaluation
of the pseudo-producing time of a smoothly varying flow rate
history in pressure buildup analyses. From pressure-transient
theory, Palacio and Blasingame6 established that during the
pseudo-steady state flow regime (fully boundary dominated
flow in a closed system), that the material balance time
function is exactly equal to the rigorous superposition-in-time
relationship for the pressure-transient solution of a varying
flow rate history.

For rate-transient analyses, we can also define a material


balance time approximation for oil reservoir analyses that is
identical in general form as the material balance time
function reported by Palacio and Blasingame6.
This
relationship is expressed mathematically in Eq. 14 for ratetransient oil reservoir analyses. In the rate-transient case, the
exact relationship between the flow rate and cumulative
production functions change with each flow regime as a
function of time.

tmb ( tn ) =

24 Np ( tn )
qo ( t n )

(14)

Employing an equivalent material balance time function


analogous to that presented in Ref. 6 for pressure-transient
analyses (instead developed for rate-transient analyses of the
production performance of gas reservoirs), we define a
material balance time function for gas reservoir analyses
that is given by Eq. 15.

tamb ( tn ) =

24000Gp ( tn )
qg ( tn )

(15)

While the material balance time function has been shown


to have a theoretical basis for the pressure-transient behavior
of a well during the pseudosteady state flow regime, it is not
correct for any other pressure-transient flow regime, and is not
correct for any rate-transient flow regime1. Other authors8,9
have also mistakenly adopted and used the material balance
time function, with only a cursory review of its applicability
and with no correction or modifications, for the analysis of the
production performance of all other flow regimes in wells
besides the pseudosteady-state flow regime.
One very important fundamental inconsistency exists in all
of the work reported in Refs. 8 through 13. The inconsistency
exhibited in those references is that they all use the material
balance time function (derived from pressure-transient theory
for only the pseudosteady state flow regime) to evaluate the
rate-transient performance of wells. All of the referenced
analyses utilize the conventional flow rate decline type curve
(rate-transient) solutions in some form to evaluate the
production behavior of oil and gas wells. As clearly
demonstrated in Ref. 1, the uncorrected material balance
time function is not correct for any rate-transient solution flow
regime, not even for fully boundary-dominated flow. The
analyses presented in this paper are internally consistent in
that they use a material balance time function derived
directly from rate-transient theory and also use the appropriate
rate-transient solutions for all of the analyses. No other
previous publication has provided a truly completely
consistent and proper analysis methodology for the analysis of
production performance data of oil and gas wells using the
material balance time function.
Agarwal et al8 have also erroneously reported that the ratetransient and pressure-transient solutions are equivalent. They
are not equivalent in any flow regime, and the proof of this
fact is documented in Ref. 1. Agarwal et al8 reported
numerical simulation results of a comparison between the

material balance time function and the equivalent


superposition-in-time function.
An example of this
comparison is given for a vertically fractured well in Fig. 1.
Figure 1 appears rather benign and innocuous at first sight,
and does appear to follow a generally linear trend. However,
replotting the same data as the ratio of material balance time
to the equivalent superposition time function, as a function of
the equivalent superposition time as in Fig. 2 gives the true
picture of what is actually happening.
The results presented in Figs. 1 and 2 were generated using
a reservoir simulator constructed with the complete, rigorous,
Laplace domain, rate-transient, analytic solution of a finiteconductivity vertical fracture in an infinite-acting reservoir14.
Bounded reservoir solutions have also been generated in this
study to verify these results and findings. These results have
also been identically duplicated with a commercial finitedifference reservoir simulator15.
As clearly presented in Ref. 1, the bounding limits of each
of the flow regimes are easily identified. The material
balance to superposition time ratio has a constant value
during the bilinear flow regime of 4/3. During the formation
linear flow regime, the ratio of the material balance time to
the superposition time reaches a constant value (which is a
maximum on the graph) of 2. Not only are these two time
functions not equivalent, the ratio between the two functions
also varies continuously over the transient history of the well.
An earlier flow regime (fracture storage or fracture linear
flow regime) also exists in the transient behavior of a
vertically fractured well but is not depicted in Figs. 1 and 2
since it (1) ends very quickly (in much less time than is
generally recorded as the first data point in production data
records), and (2) is commonly masked or distorted by
wellbore storage (only applicable for pressure-transient
solutions) if it is present. During the fracture linear flow
regime, the ratio of the material balance to the equivalent
superposition time also has a constant value of 2. Even
though this flow regime is generally not observed in
production data, the theoretical basis for the solution of the
rate-transient behavior of this flow regime is presented in the
Appendix of this paper for completeness.
A late time flow regime may also exist for all types of
wells (unfractured vertical, vertically fractured, and horizontal
wells) in closed (no flow outer boundary condition) systems,
which is also not depicted in Figs. 1 and 2. In rate-transient
analyses this flow regime is simply referred to as the fully
boundary-dominated flow regime. It occurs during the same
interval in time as the pseudo-steady state flow regime of
pressure-transient solutions, but the pressure distributions in
the reservoir during the boundary-dominated flow regime of
rate-transient solutions are completely different from those
exhibited in pressure-transient solutions. The rate-transient
behavior of oil and gas wells during the boundary-dominated
flow regime is fully presented in the Appendix of this paper
for a vertically fractured well in cylindrically and
rectangularly bounded reservoirs.
Even during the radial flow regime of unfractured vertical
wells (analogous to the pseudoradial flow regime of vertically

B. POE

SPE 77691

fractured wells), the ratio of the material balance time


function to the equivalent superposition time function achieves
a stabilized numerically determined value of about 1.08, as
clearly demonstrated in Fig. 2. Therefore, for radial (or
pseudoradial) flow analyses, an error in the time function of 8
% may be acceptable and within engineering accuracy, but
errors in the time function of as much as 200 % during the
formation linear (or pseudolinear) flow regime of vertically
fractured wells are not.
The rate-transient (flow rate versus time or cumulative
production) decline curve solutions that have been
ubiquitously used in production data analyses are the
appropriate solutions to use in most all practical cases. At
least as early as 194516, rate decline curves have been
commonly used to characterize the production performance of
oil and gas wells. Fetkovich and co-workers17,18 have greatly
expanded the use and applicability of decline curve analyses
for characterizing formation and well properties from the
production performance of oil and gas wells. Blasingame and
co-workers10-13 have also reported the development of
production analyses using decline curves that also incorporate
the use of the material balance time function.
If the proper corrections are made to the material balance
time function, a modified material balance time function can
be constructed and used to obtain an effective time function
value that is equivalent in magnitude to the rigorous
superposition time function. This type of equivalent time
function would permit the analysis of production history data
points for which the flowing pressures are not known.
Therefore, a convolution analysis of all of the production
history is performed, using the known pressure data points
where they exist in a conventional convolution analysis, and
using the modified material balance time function to
evaluate the equivalent superposition time function values that
correspond to the data points at which the pressures are not
known. This is essentially what is done in the model
construction that is described in the following section of
this paper.
Model Description
The construction of a production analysis model that
incorporates the flexibility of combining a conventional ratetransient convolution analysis for production data points with
known pressures, and the modified material balance time
concept for data points at which the pressure is unknown, into
a robust and accurate production analysis system, requires
tremendous planning and care. A production analysis system
of this type has been constructed and used, and is referred to in
this paper simply as the Pressure Optional Effective Well And
Reservoir Evaluation production analysis system.
The construction of the production analysis system
presented in this paper is most readily accomplished by
generating and storing the rate-transient decline curve
solutions for a family of well types, outer boundary
conditions, and for a range of parameter values that relate to
the model under consideration. The dependent variables that
are required in the solution are the dimensionless well flow

SPE 77691

EFFECTIVE WELL AND RESERVOIR EVALUATION WITHOUT THE NEED FOR WELL PRESSURE HISTORY

rate and cumulative production as a function of time. Rate


decline curves of this type are generated and stored for a
practical range of the independent variable values.
For unfractured vertical well rate-transient type curves, the
independent variables considered were dependent on the outer
boundary condition specified. In a closed cylindrically
bounded reservoir, the dimensionless well drainage radius

( reD ) ,

referenced to the apparent wellbore radius, is the

independent variable used to generate a family of ratetransient decline type curves. In an infinite-acting reservoir
system, the radial flow steady-state skin effect is the
independent variable used to construct the family of type
curves. The latter set is of particular importance for the case
of all well types (unfractured, fractured, and horizontal) where
no sand face flowing pressures are available at all for the
convolution analysis. The details of this procedure will be
discussed in the following section of this paper.
For vertically fractured wells in infinite-acting reservoirs,
the independent variable of interest is the dimensionless
fracture conductivity

( CfD ) .

In closed reservoirs, the

fractured well decline curves are also constructed with the


dimensionless

well

drainage

area

( AD )

as

an

independent variable.
For horizontal well decline curves, a larger number of
independent parameter values must be considered. In infiniteacting systems, the dimensionless wellbore length

( LD ) ,

vertical location in the pay zone ( ZwD ) , and well bore radius

( rwD )

are all considered. The effect of the wellbore location

has been demonstrated by Ozkan19 to have a lesser impact on


the wellbore transient behavior than the dimensionless
wellbore length and wellbore radius and may be fixed at a
constant average value (equal to approximately one half) if
limitations of array storage and interpolation are encountered.
In a finite closed reservoir there is also the dimensionless well
drainage area

( AD )

that must be included in the independent

variables considered when generating that family of


decline curves.
While the production analysis model described in this
paper presently only considers the common well types and
outer boundary conditions, the analysis methodology is
entirely general. A numerical simulation model can be used to
consider any type of well and reservoir configuration desired,
and the resulting generated rate-transient decline curves are
then used in the analysis. This is the only requirement of the
Pressure Optional Effective Well And Reservoir Evaluation
production analysis methodology described in this paper, that
the dimensionless flow rate and cumulative production
transient behavior of the particular well and reservoir
configuration under consideration can be accurately generated
and stored to be used in the decline curve analysis.

The evaluation of the ratio of the material balance time


function to the rigorous equivalent superposition-in-time
function, as a function of the equivalent superposition time, is
defined in its most fundamental form for rate-transient
analyses in Eq. 16. Note that Eq. 16 directly provides the
necessary correction for the conventional material balance
time function.

tmb ( tn ) tDmb ( tn )
QpD ( tn )
=
=
te ( tn )
tD ( tn )
qwD ( tn ) tD ( tn )

(16)

Therefore, the dimensionless time, flow rate, and


cumulative production, obtained for whatever well type and
reservoir configuration considered, are used to directly
compute the correction for the material balance time
function over the entire transient history of the well. The
modified material balance equivalent time function that is
used to effectively perform the convolution for production
data points for which the sandface pressures are unknown is
obtained by simply dividing the appropriate uncorrected
material balance time function value (given by Eqs. 14 or
15) by the correction defined with Eq. 16. Therefore, the
superposition time function value can be effectively (and
internally consistently) estimated using the material balance
time function (computed from a wells production data) and
the decline curve analysis matched well and reservoir model
dimensionless rate-transient behavior.
The actual
implementation and application of this new technology in the
model is discussed in the following section of the paper.
Implementation and Application
The implementation and application of the Pressure Optional
Effective Well And Reservoir Evaluation production analysis
methodology requires the consideration of two specific cases.
Each case must be considered separately, since each requires a
different solution procedure. In the first case, a fully
determined system results that can be directly solved at each
of the production data time levels with known sandface
flowing pressures. The second case involves a two-step or
iterative evaluation procedure in order to estimate the well and
reservoir properties using both (a) an unfractured vertical well
and infinite-acting reservoir decline curve analysis and (b) a
decline curve analysis corresponding to the actual well and
reservoir configuration of the system.
The first case considered is one in which at least one
production data point (at any point in time during the entire
production history of the well) has a known flowing sandface
pressure that corresponds to the flow rate data at that point in
time. A factor as to whether or not this case applies is also
dependent upon whether or not significant completion
pressure losses are present in the system. Since typically the
wellhead flowing pressures (or possibly bottom hole flowing
wellbore pressure measurements from permanent downhole
gauges) are the well flowing pressure data that are available,
and because the completion losses are in general formation
effective permeability (and skin effect in some models)
dependent, simultaneous solution of the sandface flowing

B. POE

pressure and the formation effective permeability and skin


effect generally require an iterative solution procedure. The
first case discussed requires that the sandface flowing pressure
for at least one point in time in the production history be
known (or that the completion losses can be assumed to be
negligible in which case the sandface flowing pressures would
be assumed to be known). If the production data set and the
well conditions do not meet these requirements, then the
analysis given in the second case that follows applies.
The second case is that in which (1) none of the sandface
flowing pressures are available for any of the production data
flow rate points, (2) one in which the sandface flowing
pressures can not be estimated directly from the bottomhole
wellbore flowing pressures, such as in the case of nonnegligible completion pressure losses, or (3) the case where
you actually have an unfractured vertical well in an infiniteacting reservoir. For any of these three conditions, an initial
analysis of the early transient (infinite-acting reservoir
response) production data on an unfractured vertical well
infinite-acting reservoir decline curve set is required. This
must be done, regardless of the actual well type. With the first
two conditions listed for the second case, this initial step is
required in order to reduce the number of unknowns in the
problem by one, typically the reservoir effective permeability
is chosen as the parameter of interest.
For the first condition of the second case, none of the
necessary sandface flowing pressures are available for a
convolution analysis. Therefore, by plotting the wells flow
rate versus its cumulative production against a dimensionless
flow rate versus dimensionless cumulative production type
curve of this type, the only unknowns in the relationships
between the dimensionless and dimensional plotting functions
on the graph is the formation effective permeability, related to
the ordinate scale dimensionless flow rate function. In this
type of analysis, only the early transient (infinite-acting
reservoir behavior) is used in determining the appropriate
decline curve match.
It is important to note that for any point on the matched
decline curve, the pressure drop (or pseudopressure drop for
gas reservoir analyses) appears in the denominator of the
dimensionless flow rate and cumulative production, the
ordinate and abscissa values, respectively. Therefore, for any
point on the decline curve, the abscissa and ordinate scale
values can be used to resolve the remaining unknown of the
problem directly related to the scales of the plotting functions
since the pressure drop term cancels out of the evaluation.
This principle is applicable for the initial infinite-acting
reservoir unfractured vertical well decline curve analysis of all
of the second case conditions. It is also important to note that
the abscissa variable (dimensionless cumulative production) in
this particular analysis is referenced to the actual wellbore
radius ( rw )

that is known, not the apparent or effective well

bore radius. The only other match variable (radial flow


steady-state skin effect) in the analysis is obtained directly
from the decline curve stem matched on the graph.

SPE 77691

For the second case and first condition, the formation


effective permeability is actually generally the only parameter
estimate that is used in subsequent computations. The
transient behavior of vertically fractured or horizontal wells
are best characterized using the specific dimensionless
parameters associated with those well types (i.e.,
CfD, LD, rwD, ZwD ) and the steady state skin effect is
generally not a good way to characterize that behavior unless
the well is actually an unfractured vertical well.
The second condition of the second case also requires an
initial analysis of the production data with a set of infiniteacting reservoir unfractured vertical well decline curves to
obtain an initial estimate of the reservoir effective
permeability so that the completion pressure losses and
corresponding sandface flowing pressures may be computed.
Once again, the reservoir effective permeability is generally
the only parameter from this analysis step that is used in the
subsequent calculations.
For the last condition of the second case (unfractured
vertical well in an infinite-acting reservoir), all of the analysis
results are used. The reservoir effective permeability and the
matched radial flow steady-state skin effect values resulting
from the analysis represent the final results for those
parameters. Once this graphical analysis step is completed,
the production data analysis is also completed for the
unfractured vertical well and infinite-acting reservoir case.
CASE 1:
The production analysis procedure that is used for the first
case discussed in this paper is accomplished in a very
straightforward manner. The dimensional flow rates of the
well are plotted against the corresponding dimensional
cumulative production at each of the production data time
levels on a log-log dimensionless flow rate versus
dimensionless cumulative production decline curve set that
corresponds to the actual reservoir type, outer boundary
condition, and well type of interest.
For each of the production data points that have known
sandface flowing pressure values, the reservoir effective
permeability is directly determined from the matched decline
curve values, the production data, and the relationship
between the dimensional and dimensionless well flow rates
(ordinate values). The system characteristic length is directly
computed from the relationship between the dimensional and
dimensionless cumulative production (abscissa values). This
means that independent estimates of these parameters can be
determined for each and every production data point for which
the sandface flowing pressures are known. While one may be
tempted to think that they may be able to evaluate how each of
these parameters change with respect to time, this is not the
case. There are two reasons why this is not possible; (1) the
convolution integral as employed in this analysis does not
permit the use of a non-linear function (reservoir model)
which would be implied if either of these parameters change
with respect to time, and (2) the rate-transient decline curve
solutions employed in the analysis have been generated for

SPE 77691

EFFECTIVE WELL AND RESERVOIR EVALUATION WITHOUT THE NEED FOR WELL PRESSURE HISTORY

constant system properties.


Therefore, these are just
independent estimates of the two parameters determined for
each of the known sandface flowing pressure data points and
should simply be averaged to report representative values for
these parameters. Statistical analysis techniques have also
been employed in the averaging process to minimize the
effects of outliers in the computed results for these parameters.
With the reservoir effective permeability and system
characteristic length known from the analysis just described,
the other well and reservoir properties are subsequently
determined from the dimensionless parameters associated with
the matched dimensionless solution decline curve stem. For
example, an unfractured well in a closed cylindrically bounded
reservoir has decline curve stems that are associated with the
dimensionless well drainage radius, referenced to the system
characteristic length. Therefore, the wells effective drainage
radius and drainage area can be readily computed from the
match result. The radial flow steady-state skin effect is also
directly obtained from the matched system characteristic
length and the wellbore radius using the effective wellbore
radius concept.
It should be noted that for closed finite reservoir decline
curve analyses, the decline curve sets displayed on the graphs
that are used for matching purposes should be modified using
the appropriate pseudosteady state coupling relationship for
the well model of interest, analogous to the systematic
methodology proposed in Ref. 13. This results in all of the
boundary-dominated flow regime decline data of the decline
curves in the set collapsing to a single decline stem on the
displayed graph, which greatly aids in graphical
matching efforts.
Similarly, for vertically fractured wells in closed
rectangularly bounded reservoirs the decline curve stems
correspond to specific values of the dimensionless fracture
conductivity and the dimensionless drainage area of the well.
The dimensional fracture conductivity is computed from the
matched dimensionless fracture conductivity, and the average
estimates of the reservoir effective permeability and fracture
half-length (equal to the matched system characteristic
length). The wells drainage area is directly computed from
the matched dimensionless well drainage area

( AD )

and the

system characteristic length


A similar scenario exists for the production analysis of a
horizontal well in a closed finite reservoir. In this case the
decline stems correspond to values of the dimensionless
wellbore length in the pay zone (referenced to the net pay
thickness), the dimensionless well effective drainage area, the
dimensionless well vertical location in the pay zone (if this
parameter is considered as variable in the analysis), and the
dimensionless wellbore radius. The total effective length of
the wellbore in the pay zone may be computed as an average
of twice the matched system characteristic length and the
value of effective wellbore length derived from the matched
dimensionless wellbore length and the net pay thickness. The
effective wellbore radius is computed from the matched
dimensionless wellbore radius and the net pay thickness. The

wells effective drainage area is readily obtained from the


matched dimensionless drainage area and the system
characteristic length.
CASE 2:
The analysis procedure that is used for the second case
discussed in this paper requires a two-step or iterative solution
procedure. As previously mentioned, the initial analysis step
for the second case involves matching the early transient data
(infinite-acting reservoir behavior) of the actual well on an
infinite-acting reservoir unfractured vertical well decline curve
set. This step is used to determine an initial estimate of the
formation effective permeability.
For the first condition of the second case discussed, where
none of the flowing pressures are known in the production
history, this is the only practical way of reliably estimating the
reservoir effective permeability independently from the effects
of all other parameters governing the rate-transient response of
the system. If this condition of the second case is applicable
in the production analysis, only estimates of the well and
reservoir properties can be obtained from the analysis since all
of the subsequent computations for the other parameter
estimates are dependent on the accuracy of the reservoir
effective permeability estimate obtained in this step.
This point may appear to be of minor significance,
however, consider the case of a vertically fractured well that
has only exhibited bilinear or pseudolinear flow in the
production data record. In both of these flow regimes (in fact
in all of the transient behavior prior to the onset of
pseudoradial flow), the apparent radial flow skin effect
exhibited by the system is transient, i.e. changes continuously
with respect to time. Only once the pseudoradial flow regime
has been exhibited in the transient behavior of the well has the
flux distribution in the fracture stabilized such that the
transient behavior of the vertically fractured well can be
characterized by a meaningful and constant steady-state radial
flow apparent skin effect. Prior to that point in time, the
production data decline on the graph may not exactly follow a
single transient decline stem, which has an associated constant
radial flow skin effect value. However, in spite of this
limitation, it has been found empirically by matching
numerous sets of numerical simulation transient production
results of fractured wells that the error resulting from inverting
the production data record in this manner to estimate the
reservoir effective permeability is generally small, typically
less than 5 %.
Since the early transient behavior of low dimensionless
conductivity (CfD < 10) vertical fractures may not generally
follow a single constant skin effect decline stem on the
unfractured vertical well and infinite-acting reservoir decline
analysis graph, it would also not be expected that the skin
effect derived from the analysis would be appropriate for
characterizing the transient behavior of the well. For higher
dimensionless conductivity (CfD > 50) fractures, the early
transient production decline data do generally tend to follow a
single decline stem. However, in general only the estimate of
the reservoir effective permeability is used in the subsequent

analyses of the production data and the remaining well and


reservoir specific parameters of interest are obtained using a
decline curve analysis that corresponds to those particular well
and reservoir conditions. A similar discussion could also be
given for the early transient behavior of horizontal wells, with
their model specific early transient flow regimes, in which
case the reservoir effective permeability is also the only
parameter estimate retained from the initial unfractured
vertical well and infinite-acting reservoir decline
curve analysis.
Once the reservoir effective permeability has been
estimated from the initial analysis step previously discussed,
the production data are then plotted on a decline curve set for
the actual well and reservoir conditions of interest. With the
previously determined reservoir effective permeability
estimate, the only unknown remaining unresolved between the
dimensionless parameter scales of the reference decline curve
set and the dimensional production data is the system
characteristic length, which is associated with the abscissa
scale of each of the matched production data points.
As previously discussed in this paper, at each production
data point on the matched decline curve stem of the graph, the
pressure (or pseudopressure) drop terms present in the
definitions of both the dimensionless flow rate and cumulative
production variables cancel out when resolving the ordinate
and abscissa match points of the dimensionless and
dimensional scales for each of the matched points. Therefore,
independent estimates of the system characteristic are directly
evaluated for each of the wells actual production data flow
rate points. Also, as previously discussed in this paper, a
statistical analysis of the independent estimates of the system
characteristic length is employed to obtain a representative
average value for this parameter.
With estimates of the reservoir effective permeability and
system characteristic length obtained in the manner just
described, the remaining unknowns of the decline curve
production analysis are obtained in the same manner as
previously given for the first case. For the third condition of
the second case, where the well is actually an unfractured
vertical well and the reservoir is still infinite-acting at the end
of the historical production data record, the analysis may be
repeated using the infinite-acting reservoir unfractured well
decline curve set to improve the estimates of the reservoir
effective permeability and steady state skin effect.
For the first and second conditions of the second case
discussed, an iterative procedure is used to update the
parameter estimates used in the completion loss and sand face
pressure calculations, whether these are measured values
(condition 2) or synthetic values (conditions 1 and 2) as
detailed in the following discussion. The iterative matching
process for this case and these conditions uses a reference
dimensionless decline curve set that corresponds to the actual
well and reservoir type being considered. The iterative
matching and analysis process are continued until convergence
and a satisfactory decline analysis match are achieved.
Concurrently with the graphical analysis matching, the
sand face flowing pressure history of the well is synthetically

B. POE

SPE 77691

generated in a systematic point-by-point manner (beginning


with the initial production data point) by resolution of the
matched dimensionless decline curve stem solution (and the
corresponding dimensionless time scale associated with that
curve) and the superposition relationships given in Eqs. 4 and
5. Definitions of the dimensionless variables used in these
relationships have been given previously in Eqs. 6 through 13.
Note that this solution procedure for estimating the sand
face flowing pressures at each of the production data flow rate
points is applicable to all well and reservoir types, and can be
performed whether any historical measured well flowing
pressures are available or not. If some sand face pressures are
known (such as in the first case discussed), a direct
comparison of the actual and synthetic sand face flowing
pressure values can be used to verify the quality of the decline
curve analysis match obtained for the production data set. The
well bore bottom hole flowing pressures can also be backcalculated from the synthetic sand face flowing pressure
history by including the completion losses of the system5.
Field Examples and Discussion
The use of the production analysis model described in this
paper is best demonstrated with some illustrative field
examples. Numerous synthetic examples have been evaluated
in the testing and validation of the model, however, field
examples have been chosen for the paper since they provide
an additional complexity in the analysis due to the fact that the
production performance of the wells are often not recorded
under the most ideal of conditions in the field. In this paper,
two examples have been chosen, which demonstrate some of
the advantages and capabilities of this production analysis
technique, for which independent estimates of the well and
reservoir properties are available. The independent estimates
of these properties are derived from conventional production
analyses or geophysical measurements such as core analyses.
The first example selected is that of a vertically fractured
gas well located in South Texas for which a complete flowing
tubing pressure record is available, which permits a
conventional convolution analysis of the production
performance of the well to evaluate the well and reservoir
properties. The second example is that of an unfractured
vertical well completed in a heavy oil reservoir in South
America (produced with an electrical submersible pump (ESP)
for which there are no pump intake pressures recorded) that
has a fairly complete set of laboratory core analyses from
whole cores.
Figure 3 is the decline curve match of the first example
well that had been previously analyzed with a production
analysis history matching model to obtain estimates of the
reservoir effective permeability, fracture half-length, and
conductivity of 0.05 md, 80 ft, and 0.5 md-ft, respectively. It
can be seen that the production analysis model discussed in
this paper provided essentially the same results (kg=0.049 md,
Xf=83 ft, kfbf=0.41md-ft) as the production analysis that used
the conventional rate-transient convolution analysis.
The oil well example production analysis required the twostep decline analysis of the production data. Figure 4 is the

SPE 77691

EFFECTIVE WELL AND RESERVOIR EVALUATION WITHOUT THE NEED FOR WELL PRESSURE HISTORY

decline curve analysis of the early transient (infinite-acting


reservoir) production performance of the well used to
determine the estimate of the reservoir effective permeability.
The production analysis resulted in an estimate of the average
reservoir effective permeability for the 54 ft thick sand of 1.28
md that is in excellent agreement with the average
permeability reported from the core analyses of 1.4 md.
Therefore, the production data analysis methodology reported
in this paper makes it both possible and practical to reliably
estimate the insitu reservoir effective permeability from the
production behavior of a well with absolutely no measured
well flowing pressures with which to perform a conventional
convolution analysis of the production performance of
the well.
The final step in the oil well example decline curve
analysis is depicted in Fig. 5. This graph is an illustration of
the decline analysis matching performed to evaluate the radial
flow steady state skin effect and obtain an estimate of the
effective well drainage area. There is not an independent
estimate of the steady state skin effect available for the well
with which to compare but the inverted estimate of skin effect
is consistent with the wells completion type and performance.
The effective well drainage area estimate obtained from the
analysis of 194 acres is also in good agreement with the well
spacing of about 200 acres on which the wells in this field
have been drilled.
Additional examples that demonstrate the application of
this powerful production data analysis methodology could also
be readily presented. However, some of the more subtle
advantages of this production analysis technique are probably
best presented in a discussion format.
As an example, the production analysis technique
presented in this paper of matching the well flow rate versus
the wells cumulative production, against a log-log graph of
the decline curve data set consisting of the dimensionless well
flow rate versus the dimensionless cumulative production,
besides having the advantage of not requiring that the sand
face flowing pressures be known for each of the production
data points plotted on the graph, also directly eliminates most
all of the problems encountered in conventional convolution
analyses related to partial day or partial month production in
the production data record. If the well is only on production
for part of a day (or month if monthly production data are
used), it is often not readily apparent how to choose an
average flowing pressure to assign to that production data
point and time value in the conventional convolution analysis.
Most importantly of all is that with this production analysis
technique, values of the well flowing pressure need not be
guessed or estimated for the missing pressure values to
complete the convolution analysis of the production data. It is
also readily apparent from the theory provided in the
Appendix of this paper and in Ref. 1, as well as the oil well
ESP example in this paper, that this production analysis
technique results in an effectively rigorous convolution
analysis of the production data, even with no sand face
flowing pressures to use in the production data analysis.

Another example of a commonly encountered production


analysis problem is that of the case where an operator has
acquired the producing well from another operator that did not
record the flowing tubing pressure history of the well. The
original operator may not have even made it a practice to
maintain their own database of the production flow rate and
cumulative production history of their wells and simply rely
on a commercial production data service to keep track of the
production records of the well. Thus the current operator is
left with the daunting task of trying to evaluate the current
well condition and reservoir properties with no practical or
mathematically correct way of accomplishing that task using
the conventional convolution techniques currently employed
in valid production data analysis models.
A similar scenario arises in the case of an oil well that
would initially produce naturally on its own, but later required
the use of an artificial lift system in order to continue
producing. While the well was initially flowing under its own
reservoir drive energy, the flowing wellhead pressures could
readily be recorded and used in the convolution analysis.
However, once the well was switched over to a pump (either
ESP or beam pump) the wellhead flowing pressure would not
generally be recorded or would be meaningless in the analysis
if it were. In this case, if an evaluation of the current well
condition is required, the analyst is faced with the problem of
how to perform the convolution analysis of the production
history while the well was being artificially lifted.
The production analysis techniques reported in this paper
provide for the first time a truly mathematically correct,
internally-consistent, and practical means of effectively
performing a convolution analysis of these types of production
analysis problems to permit the estimation of the well and
reservoir properties.
Conclusions
The major conclusions that can be drawn from this work are
the following:
1. An effective convolution analysis of a wells production
performance that has a partial or absent flowing pressure
record can be accomplished using a corrected material
balance time function and the rate-transient solutions
using the analysis procedures presented in this paper.
2. The uncorrected material balance time function is only
truly applicable for the pressure-transient analysis of a
wells production behavior during the pseudosteady state
flow regime. It is not truly correct for any other pressuretransient flow regime and is incorrect for all rate-transient
flow regimes of all well types.
3. The production analysis techniques presented in this paper
permit for the first time a mathematically correct means of
evaluating the production performance of artificially lifted
wells to obtain estimates of their well and
reservoir properties.
4. The effective convolution analysis reported in this paper is
entirely general and is applicable to all well conditions and
reservoir types as long as the appropriate dimensionless
rate-transient solution can be generated for those cases.

10

B. POE

Nomenclature
A Well drainage area, ft2

AD
bf
Bo
CfD
ct
ctf
fBF

Dimensionless drainage area, AD = A / LC


2

Fracture width, ft
Oil formation volume factor, rb/STB
Dimensionless fracture conductivity, CfD = kfbf / kXf
Reservoir total system compressibility, 1/psia
Fracture total system compressibility, 1/psia
Cumulative production bilinear flow superposition time
function
Flow rate bilinear flow superposition time function

fBF 1
fFL Cumulative production formation linear flow
fFL1
fFS
fFS 1
Gp
h
kf
kg
ko
LC
LD
Lh
m
n
Np
pD
pDi
pi
pp
psc
pwD
pwf
qD
qg
qo
QpD
qwD
re
reD
rw

superposition time function


Flow rate formation linear flow superposition time
function
Cumulative production fracture storage linear flow
superposition time function
Flow rate fracture storage linear flow superposition
time function
Cumulative gas production, MMscf
Reservoir net pay thickness, ft
Fracture permeability, md
Reservoir effective permeability to gas, md
Reservoir effective permeability to oil, md
System characteristic length, ft
Dimensionless horizontal well length in pay zone,

LD = Lh / 2h

Effective horizontal well length in pay zone, ft


Summation index
Index of current or last data point
Cumulative oil production, STB
Dimensionless pressure solution
Dimensionless pressure at the ith time level
Initial reservoir pressure, psia
Real gas pseudopressure potential, psia2/cp
Standard condition pressure, psia
Dimensionless well bore pressure
Sand face flowing pressure, psia
Dimensionless flow rate
Gas flow rate, Mscf/D
Oil flow rate, STB/D
Dimensionless cumulative production
Dimensionless well flow rate
Effective well drainage radius, ft
Dimensionless well drainage radius, reD = re / LC
Well bore radius, ft

SPE 77691

rwD
T
ta
tae
tamb
tD
tDi
te
ti
tmb
tn
TSC
XD
XD *
XeD
Xf
XwD
YD
YeD
YwD
ZwD

Dimensionless well bore radius, rwD = rw / h


Reservoir temperature, deg R
Pseudotime integral transformation, hr-psia/cp
Equivalent pseudotime superposition function,
hr-psia/cp
Gas reservoir material balance time, hr
Dimensionless time
ith dimensionless time in production history
Equivalent time superposition function, hr
ith time level in production history, hr
Oil reservoir material balance time, hr
Last or current time level in production history, hr
Standard condition temperature, deg R
Dimensionless X direction spatial position
Dimensionless fracture spatial position
Dimensionless X direction drainage areal extent
Effective fracture half-length, ft
Dimensionless X direction well spatial position
Dimensionless Y direction spatial position
Dimensionless Y direction drainage areal extent
Dimensionless Y direction well spatial position
Dimensionless well vertical spatial position

Greek

fD

Dimensionless parameter
Dimensionless parameter
Reservoir effective porosity, fraction BV
Fracture effective porosity, fraction BV
Pseudoskin due to dimensionless fracture conductivity
Pseudoskin due to bounded nature of reservoir
Dimensionless fracture hydraulic diffusivity

gct Mean value gas viscosity-total system compressibility,


o

cp/psia
Oil viscosity, cp

Functions
erfc Complimentary error function
exp Exponential function

ln

Natural logarithmic function

Acknowledgements
The author would like to express his appreciation to
Schlumberger for the permission to publish the results of this
research effort.

SPE 77691

EFFECTIVE WELL AND RESERVOIR EVALUATION WITHOUT THE NEED FOR WELL PRESSURE HISTORY

References
1. Poe, B.D. Jr.: Oil and Gas Reservoir Production Analysis
Apparatus and Method With an Effective Convolution
Analysis That Does Not Require Well Pressure History,
U.S. Patent pending.
2. van Everdingen, A.F. and Hurst, W.: The Application of
the Laplace Transformation to Flow Problems in
Reservoirs, Trans., AIME (1949) 186, 305-324.
3. Poe, B.D. Jr., Conger, J.G., Farkas, R., Jones, B., Lee,
K.K., and Boney, C.L.: Advanced Fractured Well
Diagnostics for Production Data Analysis, paper SPE
56750 presented at the 1999 Annual Technical Conference
and Exhibition, Houston, TX, Oct. 3-6.
4. Poe, B.D. Jr. and Marhaendrajana, T.: Investigation of the
Relationship Between the Dimensionless and Dimensional
Analytic Transient Well Performance Solutions in LowPermeability Gas Reservoirs, paper SPE 77467 presented
at the 2002 SPE Annual Technical Conference and
Exhibition, San Antonio, TX, Sept. 29 Oct. 2.
5. The Technology of Artificial Lift Methods, Brown, K.E.
(ed.), 4 PennWell Publishing Co., Tulsa, OK (1984).
6. Palacio, J.C. and Blasingame, T.A.: Decline-Curve
Analysis Using Type Curves Analysis of Gas Well
Production Data, paper SPE 25909 presented at the 1993
SPE Rocky Mountain Regional / Low Permeability
Reservoirs Symposium, Denver, CO, Apr. 12-14.
7. McCray, T.L.: Reservoir Analysis Using Production
Decline Data and Adjusted Time, M.S. Thesis, Texas
A&M University, College Station, TX (1990).
8. Agarwal, R.G., Gardner, D.C., Kleinsteiber, S.W., and
Fussell, D.D.: Analyzing Well Production Data Using
Combined Type Curve and Decline Curve Analysis
Concepts, SPE Res. Eval. and Eng., (Oct. 1999) Vol. 2,
No. 5, 478-486.
9. Crafton, J.W.: Oil and Gas Well Evaluation Using the
Reciprocal Productivity Index Method, paper SPE 37409
presented at the 1997 Production Operations Symposium,
Oklahoma City, OK, Mar. 9-11.
10. Doublet, L.E., Pande, P.K., McCollum, T.J., and
Blasingame, T.A.: Decline Curve Analysis Using Type
Curves Analysis of Oil Well Production Data Using
Material Balance Time: Application to Field Cases, paper
SPE 28688 presented at the 1994 SPE Petroleum
Conference and Exhibition of Mexico, Veracruz, MX,
Oct. 10-13.
11. Shih, M.Y. and Blasingame, T.A.: Decline Curve
Analysis Using Type Curves: Horizontal Wells, paper
SPE 29572 presented at the 1995 SPE Joint Rocky
Mountain Regional and Low Permeability Reservoirs
Symposium, Denver, CO, Mar. 19-22.
12. Doublet, L.E. and Blasingame, T.A.: Decline Curve
Analysis Using Type Curves: Water Influx/Waterflood
Cases, paper SPE 30774 presented at the 1995 SPE
Annual Technical Conference and Exhibition, Dallas, TX,
Oct. 22-25.
13. Doublet, L.E. and Blasingame, T.A.: Evaluation of
Injection Well Performance Using Decline Type Curves,

11

paper SPE 35205 presented at the 1995 SPE Permian Basin


Oil and Gas Recovery Conference, Midland, TX,
Mar. 27-29.
14. Poe, B.D. Jr., Shah, P.C., and Elbel, J.L.: Pressure
Transient Behavior of a Finite-Conductivity Fractured
Well With Spatially Varying Fracture Properties, paper
SPE 24707 presented at the 1992 SPE Annual Technical
Conference and Exhibition, Washington D.C., Oct. 4-7.
15. SABRETM - A General Purpose Petroleum Reservoir
Simulator, S.A. Holditch & Associates, Inc., College
Station, TX, Oct. 1993.
16. Arps, J.J.: Analysis of Decline Curves, Trans., AIME
(1945) 160, 228-247.
17. Fetkovich, M.J. Decline Curve Analysis Using Type
Curves, JPT (June 1980) 1065-1077.
18. Fetkovich, M.J. et al: Decline Curve Analysis Using Type
Curves Case Histories, SPEFE (Dec. 1987) 637-656.
19. Ozkan, E.: Performance of Horizontal Wells, Ph.D.
dissertation, University of Tulsa, Tulsa, OK (1988).
20. Stehfest, H.: Numerical Inversion of Laplace
Transforms, Communications of the ACM (Jan. 1970), 13,
No. 1, 47-49 (Algorithm 368 with corrections).
Appendix
Relationship Between Material Balance Time Function
and Superposition Time Function for Vertically
Fractured Wells
Specialized analyses can be derived for oil and gas reservoirs
for each of the flow regimes of a vertically fractured well.
The flow regimes that will be considered in this discussion are
the rate-transient flow regimes of a fractured well that
correspond to the fracture storage linear, bilinear, formation
linear or pseudolinear, pseudoradial, and boundary-dominated
flow regimes.
Fracture Storage Linear Flow Regime
The wellbore dimensionless rate transient behavior of a
fractured well during the fracture storage linear flow regime is
given3 by Eq. A-1.

qwD ( tD ) =

CfD
fD tD

= 0.179587

CfD
fD tD

(A-1)

The dimensionless cumulative production of a fractured


well during the fracture storage linear flow regime is given3 by
Eq. A-2.
1/ 2

2CfD tD
QpD ( tD ) = 3/ 2

fD

1/ 2

tD
= 0.359174CfD

fD

(A-2)

The dimensional well bore rate-transient behavior of a


fractured oil well during the fracture storage linear flow
regime is presented in Eq. A-3 and the corresponding
superposition-in-time function for this flow regime is given in
Eq. A-4.

qo ( tn )
h bf kf f ctf
fFS 1 ( tn )
=
pi pwf ( tn ) 12.7677 Bo o

(A-3)

12

B. POE

n 1

fFS 1 ( tn ) =
i =1
n >1

pi pwf ( ti )

pi pwf ( tn ) tn ti 1

tn ti
tn tn 1
1

(A-4)
The pressure drop normalized cumulative production of a
fractured oil well during the fracture storage linear flow
regime is given by Eq. A-5 and the corresponding
superposition-in-time function for this flow regime is given in
Eq. A-6.

N p ( tn )
h bf kf f ctf
=
fFS ( tn )
pi pwf ( tn ) 153.212 Bo o
n 1

fFS ( tn ) =
i =1
n >1

(A-5)

tn ti 1 tn ti + tn tn 1
pi pwf ( tn )

(A-6)
The superposition-in-time functions are directly related to
one another during the fracture storage linear flow regime of a
fractured oil well as expressed in Eq. A-7.

fFS 1 ( tn )

(A-7)

The relationship between the material balance time


function and the rigorous superposition time function for the
fracture storage linear flow regime behavior of fractured oil
wells is given by Eq. A-8.

tmb ( tn ) = 2 te ( tn )

(A-8)

The dimensional well bore rate-transient behavior of a


fractured gas well during the fracture storage linear flow
regime is presented in Eq. A-9.

qg ( tn )

pp ( pi ) pp ( pwf ( tn ) )

hTscbf kf f ctf
4548.06 pscT ct ( tn )

fFS1 ( ta ( tn ) )
(A-9)

n 1

fFS 1 ( ta ( tn ) ) =
i =1
n >1

pp ( pi ) pp ( pwf ( ti ) )
1

pp ( pi ) pp ( pwf ( tn ) ) ta ( tn ) ta ( ti 1)

1
+

ta ( tn ) ta ( tn 1 )
ta ( tn ) ta ( ti )
1

(A-10)
Eq. A-11 gives the corresponding pseudopressure drop
normalized cumulative production behavior of a fractured well
during the fracture storage linear flow regime.
Gp ( tn )
pp ( pi ) pp ( pwf ( tn ) )

hTscbf

g ( tn )

gct ( tn )

54577300 psc T

n 1

fFS ( ta ( tn ) ) =
i =1
n >1

pp ( pi ) pp ( pwf ( ti ) )

t a ( t n ) ta ( ti 1 )
pp ( pi ) pp ( pwf ( tn ) )

ta ( tn ) ta ( ti ) + ta ( tn ) ta ( tn 1)

(A-12)
The relationship between the flow rate and cumulative
production superposition-in-time functions for a fractured gas
well during the fracture storage linear flow regime is given by
Eq. A-13 and is equal to the equivalent pseudotime function
for a varying sandface pressure history.

tae ( tn ) = fFS ( ta ( tn ) ) =

pi pwf ( tn )

te ( tn ) = fFS ( tn ) =

SPE 77691

kf fcft

fFS ( ta ( tn ) )

(A-11)

fFS 1 ( ta ( tn ) )

(A-13)

The proof of the identities established in Eqs. A-7 and A13 for fractured oil and gas well behaviors during the fracture
storage linear flow regime can be accomplished by at least one
of four possible means: (1) by simple numerical evaluation,
(2) function (series) analysis, (3) considering a simple case
such as a single inner boundary condition value production
history, or (4) simply proven heuristically. The first two
options are rather tedious and will not be presented in this
discussion. The third method of proof is trivial, since for n=1
the two functions are easily proven to be exactly equal.
Heuristically we know that for any given value of dimensional
time during the fracture storage linear flow regime that the
flow rate and cumulative production superposition-in-time
functions must relate directly to the same value of
dimensionless time. This general and fundamental heuristic
principle applies to all flow regimes of all types of wells.
Therefore, the relationship between the material balance
time and superposition time functions for the fracture storage
linear flow regime of a fractured gas well is given by
Eq. A-14.

tamb ( tn ) = 2 g ct ( tn ) tae ( tn )

(A-14)

Bilinear Flow Regime


The wellbore dimensionless rate-transient behavior of a
fractured well during the bilinear flow regime is given3 by
Eq. A-15.

qwD ( tD ) =

2 CfD
CfD
= 0.367351 1/ 4
tD
3
tD1/ 4
4

(A-15)

The corresponding dimensionless cumulative production of


a fractured well during the bilinear flow regime is given by
Eq. A-16.

QpD ( tD ) =

4 2 CfD 3/ 4
tD = 0.489801 CfD tD 3/ 4 (A-16)
3
3
4

The dimensional rate-transient behavior of a fractured oil


well during the bilinear flow regime is given by Eq. A-17.

SPE 77691

EFFECTIVE WELL AND RESERVOIR EVALUATION WITHOUT THE NEED FOR WELL PRESSURE HISTORY

qo ( tn )
h kfbf ko ct
=

pi pwf ( tn ) 48.9821Bo o3

1/ 4

(A-17)

pi pwf ( ti )
1
1

1/ 4
1/ 4
pi pwf ( tn ) ( tn ti 1)
( tn ti )

n 1

fBF 1 ( tn ) =
i =1
n >1

fBF 1 ( tn )

1
1/ 4

N p ( tn )
h kfbf ko ct
=

pi pwf ( tn ) 881.678 Bo o 3

1/ 4

n 1

i =1
n >1

fBF ( tn )

(A-19)

pi pwf ( ti )
( tn ti 1)3/ 4 ( tn ti )3/ 4

pi pwf ( tn )

+ ( tn tn 1 )

3/ 4

(A-20)

The equivalent time function that can be used for the ratetransient behavior of a fractured oil well during the bilinear
flow regime is given by Eq. A-21.

te ( tn ) = fBF ( tn )

4/3

1
fBF 1 ( tn )

4
tmb ( tn ) = te ( tn )
3

(A-22)

The dimensional rate-transient behavior of a fractured gas


well during the bilinear flow regime is given by Eq. A-23.

qg ( tn )

hTsc ( kg )
kfbf
=
fBF ( ta ( tn ) )
17448.2 psc T
pp ( pi ) pp ( pwf ( tn ) )
1/ 4

(A-23)

fBF 1 ( ta ( tn ) ) =
i 1
n >1

pp ( pi ) pp ( pwf ( ti ) )
1

1/ 4
pp ( pi ) pp ( pwf ( tn ) ) ( ta ( tn ) ta ( ti 1 ) )
1

(t (t ) t (t ))

1/ 4

fBF ( ta ( tn ) ) =
i =1
n >1

1
+
1/ 4
( ta ( tn ) ta ( tn 1 ) )
(A-24)

The dimensional pseudopressure drop normalized


cumulative production behavior of a fractured gas well during
the bilinear flow regime is given by Eq. A-25.

(A-25)

pp ( pi ) pp ( pwf ( ti ) )

( ta ( tn ) ta ( ti 1) )3/ 4
pp ( pi ) pp ( pwf ( tn ) )

( t a ( t n ) t a ( ti ) )

3/ 4

+ ( ta ( tn ) ta ( tn 1) )3/ 4

(A-26)
The relationship between the flow rate and cumulative
production superposition time functions for a fractured gas
well during the bilinear flow regime is given by Eq. A-27.
This relationship also constitutes the equivalent pseudotime
that can be used in the rate-transient analysis of the behavior
of a fractured gas well during the bilinear flow regime.

tae ( tn ) = fBF ( ta ( tn ) )

4/3

1
fBF 1 ( ta ( tn ) )

(A-27)

Therefore, the relationship between the material balance


and the superposition time function for the bilinear flow
regime of a fractured gas well is given by Eq. A-28.

4
g ct ( tn ) tae ( tn )
3

tamb ( tn ) =

(A-21)

The rate-transient equivalent material balance time


approximation for the bilinear flow regime of a fractured oil
well is therefore given by Eq. A-22.

n 1

n 1

hTsc ( kg )
kfbf
fBF ( ta ( tn ) )
314071000 psc T
1/ 4

pp ( pi ) pp ( pwf ( tn ) )

(A-18)

( tn tn 1 )

The pressure drop normalized cumulative production


behavior of a fractured oil well during the bilinear flow regime
is given by Eq. A-19.

fBF ( tn ) =

Gp ( tn )

13

(A-28)

Formation Linear Flow Regime


The dimensionless rate-transient behavior of a high
conductivity (CfD > 300) fractured well during the formation
linear flow regime is given3 by Eq. A-29.

qwD ( tD ) =

tD

0.359174
tD

(A-29)

The dimensionless cumulative production of a fractured


well during the formation linear flow regime is given3 by Eq.
A-30.

QpD ( tD ) =

4 tD
= 0.718348 tD
3/ 2

(A-30)

Eq. A-31 gives the dimensional rate-transient behavior of a


fractured oil well in the formation linear flow regime.

qo ( tn )
h Xf ko ct
=
fFL1 ( tn )
pi pwf ( tn ) 6.38385 Bo o
n 1

fFL1 ( tn ) =
i =1
n >1

(A-31)

pi pwf ( ti )

1
1
1

pi pwf ( tn ) tn ti 1
tn ti
tn tn 1

(A-32)
The dimensional pressure drop normalized cumulative
production behavior of a fractured oil well during the
formation linear flow regime is given by Eq. A-33.

14

B. POE

N p ( tn )
h Xf ko ct
=
fFL ( tn )
pi pwf ( tn ) 76.6062 Bo o
n 1

fFL ( tn ) =
i =1
n >1

(A-33)

pi pwf ( ti )

(A-34)
The relationship between the flow rate and cumulative
production superposition-in-time functions of a fractured oil
well during the formation linear flow regime is given by Eq.
A-35. This relationship also represents the equivalent time
that can be used in rate-transient analyses of fractured oil
wells during the formation linear flow regime.

fFL1 ( tn )

(A-35)

The equivalent rate-transient material balance time


function for a fractured oil well during the formation linear
flow regime is given by Eq. A-36.

tmb ( tn ) = 2 te ( tn )

(A-36)

The dimensional rate-transient behavior of a fractured gas


well during the formation linear flow regime is defined as
given in Eq. A-37.

qg ( tn )

pp ( pi ) pp ( pwf ( tn ) )

h Tsc Xf kg
fFL1 ( ta ( tn ) )
2274.03 psc T
(A-37)

n 1

fFL1 ( ta ( tn ) ) =
i =1
n >1

pp ( pi ) pp ( pwf ( ti ) )

pp ( pi ) pp ( pwf ( tn ) ) ta ( tn ) ta ( ti 1 )


a i

1
1
+

t t t t
ta tn ta tn 1

a n

pp ( pi ) pp ( pwf ( tn ) )
n 1

fFL ( ta ( tn ) ) =
i =1
n >1

tae ( tn ) = fFL ( ta ( tn ) ) =

h Tsc Xf gct ( tn ) kg
27288600 psc T

pp ( pi ) pp ( pwf ( ti ) )

fFL1 ( ta ( tn ) )

fFL ( ta ( tn ) )
(A-39)

ta ( tn ) ta ( ti 1 )

pp ( pi ) pp ( pwf ( tn ) )

ta ( tn ) ta ( ti ) + ta ( tn ) ta ( tn 1 )

(A-40)
The effective pseudotime that can be used in rate-transient
analyses of a fractured gas well during the formation linear
flow regime is presented in Eq. A-41. This expression also

(A-41)

The equivalent material balance time function that can


be derived for rate-transient analyses of fractured gas wells
during the formation linear flow regime is given by Eq. A-42.

tamb ( tn ) = 2 g ct ( tn ) tae ( tn )

(A-42)

Pseudolinear Flow Regime


The dimensionless rate-transient behavior of low to moderate
dimensionless conductivity fractured wells during the
pseudolinear flow regime is given by Eq. A-43. The
pseudolinear linear flow rate-transient solution given in Eq. A43 can be readily shown to reduce to that of Eq. A-29 for
infinite conductivity fractures.

qwD ( tD ) =

9 CfD 2 tD
3 CfD tD
3 CfD
exp
erfc

4
2

(A-43)
The rate-transient flow rate and cumulative production
behavior of vertically fractured oil and gas wells during the
pseudolinear flow regime are presented in detail in Ref. 1 and
rather lengthy. Of greater importance for this discussion
however is the resulting relationship between the material
balance time function and the equivalent superposition time
function during this flow regime for a vertically fractured
well. This result is given in Eq. A-44 for a fractured oil
well analysis.

2
1

te ( tn )

1 +

tmb ( tn ) =
exp ( ) erfc

( )

(A-38)
The pseudopressure drop normalized cumulative
production behavior of a fractured gas well during the
formation linear flow regime is given by Eq. A-39.

Gp ( tn )

provides the relationship between the flow rate and cumulative


production superposition time functions for a fractured gas
well during the formation linear flow regime.
2

tn ti 1 tn ti + tn tn 1

pi pwf ( tn )

te ( tn ) = fFL ( tn ) =

SPE 77691

(A-44)

where:

( kfbf ) te ( tn )
2

1685.55ko o ct Xf 4

(A-45)

The corresponding relationship between the material


balance and superposition time functions during the
pseudolinear flow regime of a fractured gas well is given in
Eq. A-46.

2
1

gct ( tn ) tae ( tn )

1 +

tamb ( tn ) =
exp ( ) erfc

( )

(A-46)
where:

SPE 77691

EFFECTIVE WELL AND RESERVOIR EVALUATION WITHOUT THE NEED FOR WELL PRESSURE HISTORY

( kfbf )

tae ( tn )
=
1685.55kg Xf 4

bounded reservoir during the fully developed boundarydominated flow regime may also be evaluated using Eq. A-52.
The imaging pseudoskin term due to the bounded nature of the

(A-47)

qwD ( XwD, YwD, XeD, YeD, XD, YD, tD ) =

where:

2
QpD ( tD ) = L 2

*
s 2 + ln ( 4 ) 2 + 2 ( XD , 0 ) ln ( s )

= 2

(A-50)

(A-54)
The dimensionless cumulative production of a fractured well
located in a closed rectangularly bounded reservoir during the
boundary-dominated flow regime is given by Eq. A-55.
XeDYeD
2 tD
QpD ( XwD , YwD , XeD , YeD , XD , YD , tD ) =
1 exp

2
XeDYeD
(A-55)
For both of these drainage area shapes, the relationship
between the material balance time function and the
equivalent superposition time function is given in general
form by Eq. A-56.

tmb ( tn ) 1 exp ( u )
=
te ( tn ) u exp ( u )

Boundary-Dominated Flow Regime


The dimensionless rate-transient behavior of a fractured well
that is centrally located in a closed cylindrically bounded
reservoir may be characterized3 with Eq. A-51.

[ ln ( r

eD

) + 0.25 + ( X

, 0 + XD , reD

ln ( reD ) + 0.25 + XD , 0 + XD , reD


*

(A-51)
The dimensionless cumulative production behavior of a
fractured well that is centrally located in a closed cylindrically

YeD 1 YD YD 2 + YwD 2
+

XeD 3 YeD
2YeD 2

Generally, the pseudoradial flow regime rate-transient


Laplace domain solutions (Eqs. A-48 and A-49) are inverted
into the real space domain using a numerical inversion
technique such as the Stehfest Algorithm20. The results of this
application may be found in detail in Ref. 1 and will not be
included in this discussion due to length constraints.
However, the ratio of the material balance to the equivalent
superposition time functions during the late pseudoradial flow
regime does tend to stabilize at a value of approximately equal
to 1.06 for CfD = 0.1 and 1.084 for CfD = 10000, as can be seen
in Fig. 2 of this paper.

eD

2 tD
1
exp

XeDYeD

2 XeD 1
m
m XwD
m XD
sin

cos
cos
i
2
m =1 m
XeD
XeD
XeD
YeD YD YwD + cosh m YeD ( YD + YwD )
cosh m


XeD
XeD

m YeD
sinh

XeD

(A-49)
The relationship between the material balance time and
the equivalent superposition time functions during the infiniteacting pseudoradial flow regime of a vertically fractured well
can be obtained using the fundamental relationship between
the two functions, given in Eq. A-50.

2tD

(A-53)

A-48 and A-49 is defined in Ref. 1 and is a function of the


fracture dimensionless conductivity.

exp
r

) (

(A-52)
Similarly, the boundary-dominated flow regime rate-transient
behavior of a fractured well that is located in a closed
rectangularly bounded reservoir is given by Eq. A-53.

solution. The pseudoskin term XD , 0 appearing in Eqs.

tmb ( tn ) tDmb ( tn )
QpD ( tn )
=
=
te ( tn )
tD ( tn )
qwD ( tn ) tD ( tn )

(A-48)
Similarly, the dimensionless cumulative production of a
fractured well during the pseudoradial flow regime can be
evaluated in real space using the Laplace space rate-transient
*

reD2
2tD

1
exp

*
*
2

reD ln ( reD ) + 0.25 + XD ,0 + XD , reD

*
s 2 + ln ( 4 ) 2 + 2 ( XD , 0 ) ln ( s )

QpD XD , reD , tD =

qwD ( tD ) = L1

qwD XD , reD , tD =

reservoir XD , reD is defined in Ref. 1.

Pseudoradial Flow Regime


The dimensionless rate-transient behavior of a fractured well
during the infinite-acting pseudoradial flow regime is given3
by Eq. A-48.

15

(A-56)

where:

)]

u=

2 tD
AD

(A-57)

For rectangularly bounded reservoir cases, the parameter ( )


has been defined in Eq. A-54. For a cylindrically bounded
reservoir, the appropriate value of this parameter is given by
Eq. A-58.

= ln ( reD ) + 0.25 + ( XD* , 0 ) + ( XD* , reD )

(A-58)

16

B. POE

SPE 77691

120000
100000

tmbD

80000
60000
CfD =
CfD =
CfD =
CfD =
CfD =
CfD =

40000
20000

0.1
1
10
100
1000
10000

0
0

20000

40000

60000

80000

100000

120000

tD

Fig. 1 Cartesian graph of tmbD versus tD for vertically


fractured well in infinite-acting reservoir.

Fig. 4 Vertical oil well example with ESP (no pressures),


match for oil effective permeability.

2.5
CfD
CfD
CfD
CfD
CfD
CfD

tmbD/tD

=
=
=
=
=
=

0.1
1
10
100
1000
10000

1.5

1
1.E-06

1.E-04

1.E-02

1.E+00
tD

1.E+02

1.E+04

1.E+06

Fig. 2 Semi-log graph of tmbD/tD versus tD for vertically


fractured well in infinite-acting reservoir.
Fig. 5 Vertical oil well example with ESP, evaluation of
skin effect and drainage area.

Fig. 3 Example vertically fractured gas well with known


flowing pressures.

Anda mungkin juga menyukai