Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

Dynamic Flow Analysis - v4.

02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p1/32

3 - Pressure Transient Analysis


OH – OSF - DV

The two sections following the introduction (3.A) present the classical (3.B) and modern (3.C)
methodology tools developed for Pressure Transient Analysis. In section (3.D) we will present
the typical path recommended to perform such analysis. In section (3.E) we will present how
to use the same modelling tools to design a test. The two last sections will be dedicated to
operational considerations on well testing: what data to gather (3.F) and how to validate the
gathered data (3.G).

3.A Introduction
Historically, Pressure Transient Analysis started as ‘Well Test Interpretation’, as the candidate
data for this process were only gathered during specific field operations (well tests) designed
to acquire and interpret these data. In the last 20 years the term has become increasingly
invalid, as the same processing has been applied to pressure and rate data not acquired during
a well test. Currently, the main sources of pressure transient data are well tests of various
types, formation tests and any well shut-in monitored with permanent gauges.

So today’s definition of Pressure Transient Analysis is less about the operation that produces
the data than the processing applied to this data. The principle of Pressure Transient Analysis
is the gathering of pressures and rates, preferably downhole, and the focus on a period of
interest, generally a shut-in period (build-up or fall-off) to perform a diagnostic. The diagnostic
leads to the choice of a model and this model is then used to simulate pressure values to be
matched on the period of interest. The model is then tried against a larger portion of the
recorded data, if available. Part of the process involves matching the model parameters on the
data by trying to achieve the best possible match, by trial and error or nonlinear regression.

Several items differentiate the processing of Pressure Transient Analysis from


Production Analysis:

• The selected period is preferably a period where the pressure response is clean. This is
why, with the notable exception of gas tests, the selected period will generally be a shut-in
period. Conversely Production Analysis focuses on flow (producing), though shut-ins are
not formally excluded.
• As a consequence Pressure Transient Data have the tendency to be cleaner, with a high
capacity for diagnostics compared to Production Analysis, where one will typically match a
cloud of simulated points onto a cloud of acquired data.
• The selected period is relatively short (hours, days, weeks) rather than months and years,
which is the typical time frame of Production Analysis.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p2/32

• The process consists in matching the pressures, not the rates. The rates are considered as
a major corrective function used in the calculation of the pressure derivative functions,
time scales, model convolution and data deconvolution. In Production Analysis the matched
data are generally rates, or cumulative productions, or productivity index. This being said,
Pressure Transient Analysis also allows the match on rate normalized pressures, which is
actually the inverse of the productivity index.

In the next two sections we will make a clear split with the traditional tools (the ‘old stuff’) and
modern tools (the ‘right stuff’). Traditional tools had their use in the past, and they are still in
use today, but have become largely redundant in performing analysis with today’s software.
Modern tools are at the core of today’s (2007) modern methodology, and they are of course on
the default path of the processing.

3.B The old stuff


20 years ago, the core of well test interpretation was the dual use of specialized plots and
type-curve matching:

• Specialized plots correspond to a selected scale where some flow regime of interest such as
infinite acting radial flow, linear flow, bilinear flow, spherical flow or pseudo-steady state
are characterized by a straight line. The slope and the intercept of this straight line will
generally give two parameters of the system.
• Type-curve matching consists in sliding a plot of the data, generally on a loglog scale, on
pre-printed type-curves. The relative position between the data and the type-curve, also
called time match and pressure match, gives two quantitative parameters. The choice of
type-curve will give additional information.

We will start with the semilog plots, the main specialized plots in PTA used to quantify the
main flow regime in PTA: Infinite Acting Radial Flow, or IARF.

3.B.1 IARF and Semilog plots


We have seen in Chapter 2.D that IARF is the main regime of interest in Pressure Transient
Analysis. In the case of a production at constant rate, IARF is characterized by linearity
between the pressure change and the logarithm of time. We will see that such linearity is also
found for more complex production history, provided that the right time function is used.

3.B.2 Drawdown response and MDH plot


In the case of a constant production from time 0 to infinity, we have seen in Chapter 2.D.4
that IARF is characterized, for a finite radius well in a homogeneous reservoir, by the equation:

162.6qμ ⎡ ⎛ k ⎞ ⎤
IARF equation from Chapter 2.D.4: Δp = ⎢log(Δt ) + log⎜⎜ ⎟ − 3.228 + 0.8686S ⎥
2 ⎟
kh ⎣ ⎝ Φ μct rw ⎠ ⎦
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p3/32

In the case of more complex well geometries and reservoir heterogeneities, the constant term
may be more complicated, as it will integrate the cumulative effect of these geometries and
heterogeneities. Still the response will have the same shape. The value of skin S calculated
from the equation above may not be the right value in terms of well damage according to
Darcy’s law, but it will have some meaning. It is called the ‘equivalent skin’.

The Miller-Dyes-Hutchinson (MDH) plot is a graph of the pressure or the pressure change as a
function of the logarithm of time. IARF is characterized by a linearity of the response.

Fig. 3.B.1 – Drawdown MDH plot Fig. 3.B.2 – Buildup MDH plot

Drawing a straight line through these points gives a slope and an intercept:

162.6qBμ
IARF straight line: Y = log(Δt ) + b = m log(Δt ) + b
kh
Where: b = Δp LINE (log(Δt ) = 0) = Δp LINE (Δt = 1hr )

Important: The value of b corresponds to the value of Δp on the straight line, not on the data.

Permeability and the skin are then given by:

162.6qBμ
Permeability: k=
mh

⎡ Δp ( Δt = 1hr ) ⎛ k ⎞ ⎤
Skin factor: S = 1.151⎢ LINE − log⎜⎜ ⎟
2 ⎟
+ 3.228⎥
⎣ m ⎝ Φ μct rw ⎠ ⎦

3.B.3 Build-up response and Horner plot


The MDH plot, with the simple log(∆t) time function, results directly from the log
approximation to the drawdown solution for infinite-acting radial flow. In order to use semilog
analysis for any flow period other than the first drawdown, it is necessary to take into account
superposition effects.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p4/32

We combine the build-up superposition of Chapter 2.F.2 and the drawdown IARF
approximation given in Chapter 2.D.4:

( ) (
Build-up superposition: Δp BU ( Δt ) = Δp DD t p + Δp DD (Δt ) − Δp DD t p + Δt )
162.6qBμ ⎡ ⎛ k ⎞ ⎤
IARF approximation: Δp DD ( X ) = ⎢log( X ) + log⎜⎜ ⎟ − 3.228 + 0.8686S ⎥
2 ⎟
kh ⎣ ⎝ Φ μct rw ⎠ ⎦
If Δt is large enough to reach IARF, so will tp+Δt. A lot of terms cancel out and we get:

162.6qBμ ⎛ Δt ⎞
Build-up superposition: Δp BU ( Δt ) = log⎜ ⎟ + Δp DD (t p )
kh ⎜ t + Δt ⎟
⎝ p ⎠

162.6qBμ ⎛ t p Δt ⎞ ⎡
Rearranging: Δp BU (Δt ) = log⎜ ⎟ + Δp DD (t p ) − 162.6qμ log(t p )⎤
⎜ ⎟ ⎢ ⎥⎦
kh ⎝ t p + Δt ⎠ ⎣ kh

If the production time was too short, IARF was not reached during the flow period and ∆pDD(tp)
cannot be turned into a log approximation. The constant term on the right of the equation
becomes unknown. In this case, the analysis described below will give the permeability, but
not the skin factor. If the production time was long enough, then the term tp can also be
turned into a logarithmic approximation and we get:

162.6qBμ 162.6qBμ ⎡ ⎛ k ⎞ ⎤
( )
IARF at tp: Δp DD t p − log(t p ) = ⎢log⎜⎜ ⎟
2 ⎟
− 3.228 + 0.8686 S ⎥
kh kh ⎣ ⎝ Φ μct rw ⎠ ⎦

162.6qBμ ⎡ ⎛ t p Δt ⎞ ⎛ ⎞ ⎤
So: Δp BU ( Δt ) = ⎢log⎜⎜ ⎟ + log⎜ k 2 ⎟ − 3.228 + 0.8686S ⎥ We introduce the
⎟ ⎜ Φ μc r ⎟
kh ⎢⎣ ⎝ t p + Δt ⎠ ⎝ t w ⎠ ⎥⎦
t p Δt
Horner Time as:
t p + Δt

Infinite-acting radial flow for a build-up is characterized by linearity between the pressure
response and the logarithm of Horner time. Drawing a straight line through this point gives a
slope and an intercept:
162.6qBμ ⎛ t p Δt ⎞ ⎛ t Δt ⎞
IARF straight line: Y = log⎜ ⎟ + b = m log⎜ p ⎟+b
kh ⎜ t + Δt ⎟ ⎜ t + Δt ⎟
⎝ p ⎠ ⎝ p ⎠
If the producing time tp was long enough to reach IARF, the IARF approximation for a
build-up will be similar to the drawdown relation, replacing time by Horner time, and will be
given by:

162.6qBμ ⎡ ⎛⎜ t p Δt ⎞⎟ ⎛ k ⎞ ⎤
IARF for a build-up: Δp BU ( Δt ) = ⎢log⎜ + log⎜⎜ ⎟⎟ − 3.228 + 0.8686S ⎥

⎣⎢ ⎝ t p + Δt ⎠ ⎝ Φ μct rw ⎠
2
kh ⎦⎥

After transformation, we get the equation of the Horner plot straight line in term of pressure:
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p5/32

162.6qBμ ⎛ t p + Δt ⎞
p BU = pi − log⎜⎜ ⎟⎟
kh ⎝ Δ t ⎠

Permeability-thickness product and the skin are then calculated by:

162.6qBμ
Permeability-thickness product: k =
mh

⎡ p1hr − p wf ⎛ t p +1⎞ ⎛ ⎞ ⎤
Skin factor if tp is large enough: S = 1.151⎢ + log⎜ ⎟ − log⎜ k ⎟⎟ + 3.23⎥
⎜ t ⎟ ⎜ Φμc r 2
⎢⎣ m ⎝ p ⎠ ⎝ t w ⎠ ⎥⎦

Note that the time function is such that the data plots ‘backwards’, as when ∆t is small, at the
start of the build-up, Horner function (log (tp+∆t)/∆t) will be large, and when ∆t tends to
infinite shut-in time the Horner time tends to 1, the log of which is 0.

Fig. 3.B.3 – Horner plot

If the reservoir were truly infinite, the pressure would continue to build-up in infinite-acting
radial flow and eventually intercept the y-axis at pi, the initial pressure. However, as no
reservoir is infinite, the extrapolation of the radial flow line at infinite shut-in time is called p*,
which is simply an extrapolated pressure.

If the reservoir is infinite: pi = p *

It is important to notice that the calculation of the permeability is valid even in the case of a
short production before the shut-in, while the validity of the calculation of skin is conditioned
to a producing time long enough, so IARF was reached before tp.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p6/32

3.B.4 Shut-in after a complex production and superposition plot


When a shut-in follows a complex production / injection history, the process is equivalent to
that for a simple build-up, but the Horner time must be replaced by the superposition time
function defined by:

n −1
qi − qi −1 ⎛ ⎡ n −1 ⎤ ⎞
Sn(Δt ) = ∑ log⎜ ⎢∑ Δt j − Δt ⎥ − log Δt ⎟
q n − q n −1 ⎜ j =1 ⎟
i =1 ⎝⎣ ⎦ ⎠

A plot of pressure versus superposition time is the ‘general semi-log’ plot:

Fig. 3.B.4 – General semilog: superposition plot

Calculations of permeability is the same as for the Horner plot, and the skin factor is calculated
by taking one point on the straight line, say (X,Y), and is given by:

⎧⎪Y − p wf ⎛ k ⎞ n ⎡ qi − qi −1 ⎤ ⎫⎪
S = 1.151⎨ − X − log⎜⎜ ⎟⎟ − ∑ ⎢ log(t n +1 − t i )⎥ + 3.23⎬
⎪⎩ m ⎝ Φμct rw ⎠ i =1 ⎣ q n − q n −i ⎪⎭
2

3.B.5 Complex productions and rate normalized superposition plot


In some cases the engineer will want to interpret producing periods when the rate is not
stabilized, or the rates data will be sandface rate, and therefore the value of the rate will not
instantaneously go down to zero during the shut-in period. These cases are a little more
complex than the cases above. However the principle of the semi-log analyses will be kept with
the following modifications:

For a multi-rate production or injection (whether rates are measured at surface or downhole),
the adequate semi-log plot will still have the same superposition time as the X axis, but the
pressure will be replaced by:

Q
[ pi − p(t )]
q(t )
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p7/32

where Q is a reference value for normalization, that will be generally chosen to be the last
stabilized rate of the interpreted period so that the function tends to ∆p at late time.

For a multi-rate build-up or fall-off using sandface rates, the Y axis will still be

p(t ) − p wf

but the calculation of the superposition time will allow rates to be changed during the period.
Furthermore the reference rate will not be the difference of rates before N and N-1 (this has no
meaning when dealing with continuous downhole rate measurements), but the last stabilized
rate before shut-in.

3.B.6 Loglog type-curves


3.B.6.a Origin of type-curves
We have seen in Chapter 2.C.1 that diffusion problems were solved by replacing the real
variables by dimensionless variables that eliminate other parameters in order to arrive at a set
of equations that are solved, hopefully quite simply and once and for all, in dimensionless
space. Two critical variables are the dimensionless time and dimensionless pressure.

k
Dimensionless time: t D = 0.0002637 Δt = AΔt where A = f (k , μ , rw ,...)
Φ μct rw2
kh
Dimensionless pressure: pD = Δp = BΔp where B = g ( k , h, μ ,...)
141.2qBμ

This is still used today in modern software to generate analytical models. The solution is solved
in dimensionless space, then the variables are converted to real variables and superposition is
applied to the solution, and then matched with the real data.

However, this was not possible, or at least not easy, before personal computers and related
software were available. So the ‘trick’ was to use a simple and remarkable property of the
logarithmic scales, historically used to create slide rules. Taking the logarithm of the equations
above we would get:

Logarithmic relations: log(t D ) = log(Δt ) + log( A)

and log( p D ) = log(Δp ) + log(B )

In other words, the dimensionless response, also called a type-curve, and the real response,
taken on a loglog scale, have the same shape. By translation, it is possible to match the data
on the dimensionless response and this is called a type-curve match. The value of the
translation, also called match point, in the X direction (the time match) and in the Y direction
(the pressure match) will give A and B, which in turn provides two quantitative pieces of
information. The time match will give different, model dependent, information. The pressure
match will typically give the permeability-thickness product.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p8/32

3.B.6.b Using type-curves


Type-curve matching originally consisted in plotting a loglog plot of the pressure response
versus time on tracing paper that could then be slid over a set of pre-printed type-curves, the
log cycles being square and of the same size in both plots. The selection of the type-curve
matching the data can provide one or several parameters, and the relative positions of the two
plots in the X and Y direction, also called the match point, will give two other interpretation
results.

Fig. 3.B.5 – Type-curve matching

The type-curve used for wellbore storage and skin in an infinite homogeneous reservoir (Figure
3.B.6), the pressure match (relative positions in the Y direction) will give the permeability. The
time match (relative positions in the X direction) will give the wellbore storage, and the
selection of the curve will give the skin factor.

Fig. 3.B.6 – Wellbore storage and skin type-curve


Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p9/32

3.B.7 Other specialized plot


Infinite Acting Radial flow is the most important but not the only flow regime characterized by
linearity of the pressure response when plotted versus the right time scale. Pseudo-Steady
State is another, amongst other, flow regimes of interest. The different flow regimes require a
specific plot of pressure versus the appropriate time scale for quantification of parameters
specific to the flow regime.

The main flow regimes are:

Regime Model Linearity vs


Infinite Acting Radial Flow Homogeneous Infinite logarithm of time
Pseudo-Steady State Closed systems time
Pure wellbore storage Wells with storage time
Spherical flow Limited Entry wells inverse of square root of time
Bilinear flow Fractures with Finite Conductivity fourth root of time
Linear flow Fractures square root of time
Linear flow Channel shaped reservoirs square root of time
Semi-linear flow U-shaped reservoirs square root of time

3.B.8 Equations for specialized plot other than IARF, WBS and PSS:
Equation for a linear flow (fractured well):

0.3928qB μΔt
Δp =
wh k f (φct ) f

Equation for a linear flow (Channel):

0.3928qB μΔt
Δp =
Wc h k (φct ) f

Equation for a bilinear flow (finite conductivity fracture):

37.97 qBμ
Δp = Δt 1 / 4
h Cr (φμct k )
1/ 4

Equation for a spherical flow:

70.6qBμ 2453qBμ φμct 1


Δp = −
k s rs k s3 / 2 Δt

With : (
k s = kr2 kv )
1/ 3
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p10/32

3.B.9 IPR & AOF


A common measure of the performance of any gas, or oil, well is its ability to deliver against
‘zero’ bottom hole pressure, or more accurately against atmospheric pressure. How much
could it flow with a flowing bottom hole pressure of one atmosphere? This may sound like a
pointless calculation, as it is certainly unrealistic; a gas well could not even blow out at this
rate, the bottom hole pressure would be too high. However this measurement provides an
input value for the reservoir engineer working with IPR calculations, and is an accepted
‘universal indicator’ for gas wells. It is called the ‘absolute open flow’ potential of the well,
the AOF. NB. The concept of AOF is more useful if one uses wellhead instead of bottomhole
pressures, as this indicates now the maximum achievable flowrate for that well.

In order to evaluate the AOF, the well will be tested at multiple rates. The bottom hole
pressure must be measured during each drawdown and buildup. The final drawdown is
extended, as this will provide a single ‘stabilized’ value for the calculation.

Fig. 3.B.7 – History plot

Note that frequently in well testing we ‘flow the well until stable’, although in reality the
flowing bottom hole pressure can never stabilize, except in very special circumstances. If it
did, the semi-log slope (IARF) ‘m’ would become zero, which would correspond to an infinite
kh. The pressure will always diminish, and the well is considered ‘stable’ when the pressure is
dropping very slowly. From this data the following plot is constructed, based upon the
empirical relationship that ∆m(p).vs q, on a log-log scale, will give a straight line, as long as
the m(pwf) values are recorded after the same time in each drawdown. Moreover, the ‘2-hour’,
‘4-hour’, ‘6-hour’ lines, etc., will all be parallel, converging to the ‘stabilized deliverability
curve’ for all of the stabilized points, obtained if the well had been flowed long enough for each
drawdown to stabilize:
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p11/32

Fig. 3.B.8 – AOF plot

In a typical test, as shown above, there will be only one stabilized point (Yellow), and a line is
drawn through this point parallel to the transient line. By extrapolating the stabilized
deliverability curve to a ∆m(p) value corresponding to a flowing bottom hole pressure of 14.7
psi, the AOF is obtained, indicated by the white line in Figure 3.B.7.

3.C The ‘right’ stuff


Speaking about ‘right stuff’ vs. ‘old stuff’ is deliberately provocative. Old stuff does not mean
‘wrong stuff’. What is sometimes wrong is to keep on using old techniques that turn out to be
less accurate than what was more recently developed. What is definitely wrong is to deny what
the new techniques are providing on the basis that it may be too complicated or requires
computers and not just a slide rule.

Generally these tools were very smartly designed at the time when the engineers did not have
access to modern PCs. These tools have benefited the industry in the past and were in the
development path of the modern tools we are using now. What is questionable is to continue
using such tools today, simply out of inertia, while everybody knows the limitations, and while
these limitations are successfully addressed by more modern tools generally requiring a
computer and more sophisticated processing.

3.C.1 PTA methodology before and after the 1980’s


PTA methodology, before the introduction of the derivative and the use of PCs, was essentially
a manual process alternating between type-curve matching and specialized analyses. Type-
curves without the derivative had poor diagnostic capabilities, but the result of the specialized
plots would help position the data on the type-curve. For example, drawing the IARF straight
line would give a value of permeability that could be used to set the pressure match, reducing
the uncertainties of the type-curve match. Conversely, selecting a type-curve would give the
approximate time at start of IARF, which in turn would help define the semilog plot straight
line. This process, especially in the case of gas tests, would be complemented by AOF / IPR
analysis.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p12/32

The shortcomings of this methodology were numerous

• Type-curves had very poor resolution because of the log scale. This shortcoming was later
solved by the addition of the derivative.
• Type-curves were generally suited for drawdown responses. As most of the data were
coming from build-ups, the bending of the response due to superposition was generally not
taken into account. So type-curve matching would require a very long and stable
production prior to the shut-in to make the type-curve usable.
• Type-curves were set for a discrete number of parameter values.
• Specialized plots required that a pure flow regime was established and encountered. This
would only occur in specific conditions.
• In the case of skin calculation, IARF analysis would require that the production time is long
enough; otherwise the equation giving skin would be incorrect.
• It was also very easy to draw a straight line through the two last pressure points and take
this as the IARF, even in cases where IARF was not reached.
• The process required moving back and forth between at least two different plots, if not
more.
• It was slow.

Some of the shortcomings were addressed in the early 1980’s, when Dominique Bourdet,
introduced the Bourdet Derivative. However, the derivative was also at the origin of what we
can call modern PTA methodology (hereinafter called the ‘right stuff’), extensively using model
matching with PCs, which signified the end of the methodology of the 1970’s.

To most engineers the replacement of manual techniques by computer based analysis in their
day-to-day work occurred in the 1980s, and came from three major breakthroughs:

• Electronic downhole pressure gauges, either run on memory or on electric line, became
cheap and reliable, detecting repeatable behaviours far beyond what the previous
generation of mechanical gauges could offer.
• The spread of Personal Computers allowed the development of PC-based pressure transient
analysis software. The first PC based programs appeared in the 1980’s, initially reproducing
the manual methods on a computer. Since then, new generations of tools have been
developed, with modern methodology at its core.
• In 1983 Bourdet et al published the first formulation of the pressure derivative as the slope
of the superposition semilog plot displayed on the loglog plot. The Bourdet Derivative is
certainly the single most important breakthrough in the history of Pressure Transient
Analysis. Work on differentiating the pressure response had been published before, but
Bourdet’s approach was the formulation that made it work.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p13/32

3.C.2 Bourdet derivative


The Bourdet Derivative is the slope of the semilog plot displayed on the loglog plot…

… and to be more accurate, it is the slope of this semilog plot when the time scale is the
natural log. It has to be multiplied by ln(10)=2.31 when the decimal logarithm is used in the
semilog plot.

Fig. 3.C.1 – Bourdet derivative, semilog and loglog

dΔp dΔp
For a drawdown: Δp' = = Δt
d ln (Δt ) dΔt

dΔp
For multirate: Δp ' =
d sup(Δt )

3.C.3 Bourdet derivative and early well / wellbore effects


Pure wellbore storage effects are only observed at very early time when the well pressure
behavior is dominated by the well fluid decompression (or compression).

In case of pure wellbore storage: Δp = CΔt

Even for multirate solutions at early time: sup(Δt ) ≈ ln (Δt )

dCΔt
The derivative is therefore: Δp ' = Δt = CΔt = Δp
dΔt
At early time, when pure wellbore storage is present, pressure and the Bourdet derivative
curves will merge on a unit slope straight line on the loglog plot.

Other early time flow regimes, such as linear and bilinear flow, that will be presented in more
detail later in this book, will exhibit a different and specific behavior for both pressure and the
Bourdet derivative.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p14/32

3.C.4 Early time flow regime equations


Equation for a linear flow (fractured well):

0.3928qB μΔt
Δp =
wh k f (φct ) f

The derivative is then:

0.3928qB μΔt
Δp′ = 0.5
wh k f (φct ) f

Equation for a bilinear flow (finite conductivity fracture):

37.97 qBμ
Δp = Δt 1 / 4
h Cr (φμct k )
1/ 4

The derivative is then:

37.97 qBμ
Δp ′ = 0.25 Δt 1 / 4
h Cr (φμct k )
1/ 4

Equation for a spherical flow:

70.6qBμ 2453qBμ φμct 1


Δp = −
k s rs k s3 / 2 Δt

(
With: k s = k H k v
2
)
1/ 3

The derivative is then:

⎛ 2453qBμ φμct 1 ⎞
d⎜ ⎟
⎜ Δt ⎟
Δp ' = Δt ⎝ k s3 / 2 ⎠ = −0.5 2453qBμ φμct 1
dΔt k s3 / 2 Δt

3.C.5 Bourdet derivative and IARF


When IARF occurs: Δp = m' sup(Δt )

Where m’ is the slope of the semilog straight line. In the following the drawdown response is a
specific case of the multirate response, and the logarithm of time is the specific superposition
time for a drawdown. The derivative is therefore:
dΔp
Derivative when IARF has been reached: Δp ' = = m'
d sup(Δt )
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p15/32

When IARF is reached, the derivative stabilized to a level equal to the slope of the semilog
straight line. This property was the main reason for the development of the derivative, as it
easy and straight forward to identify IARF on the loglog plot, something which is nigh on
impossible on the semilog plot. One can say that the derivative is a ‘magnifying glass’ of the
semilog behavior, conveniently placed on the same plot used, historically, for type-curve
matching.

Combined with the early time unit slope during wellbore storage, the derivative provides an
immediate way to define the pressure and the time match on the loglog plot, just by
positioning a unit slope line on the wellbore storage regime and positioning the horizontal line
on the IARF response.
This alone would have made the Bourdet derivative a key diagnostic tool. The delightful
surprise was that the derivative could do much more, and that most well, reservoir and
boundary models carry a specific signature on the derivative response. It is this remarkable
combination that allowed the derivative to become THE diagnostic and matching tool in
Pressure Transient Analysis.

3.C.6 Bourdet derivative and PSS


After a long stabilized production, when PSS is reached:

The pressure response is: Δp = AΔt + B

And the superposition time can again be approximated by sup(Δt ) ≈ ln (Δt )

d ( AΔt + B )
Δp' = Δt = AΔt
The derivative is therefore: dΔt

At very large time, we will have: Δp = AΔt + B ≈ AΔt

So, when PSS is reached, the pressure response on the loglog plot will tend to a unit slope,
while the derivative will reach this unit slope much earlier i.e.: as soon as PSS is reached, and
to ultimately merge with the pressure response. This behavior is important in Pressure
Transient Analysis but not applicable for build-ups; in a build-up the pressure stabilizes and
the derivative plunges towards zero (see the Chapter 8 on boundary effects). This behavior is
also fundamental in Production Analysis, where the main regime of interest is, precisely, the
Pseudo-Steady State, or boundary dominated flow.

3.D Modern PTA methodology


Modern Pressure Transient Analysis is based on the use of PC based PTA software products.
The key for any modern software is to combine user friendliness and a powerful technical
kernel, with both analytical and numerical capabilities. In terms of methodology, the central
diagnostic tool is the loglog plot, showing both pressure and the Bourdet derivative used for
the diagnostics and the match with the selected model(s). The section below describes the
typical path of today’s (2007) Pressure Transient Analysis. It was our understanding of what
this typical path should be that led us to implement this in Saphir.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p16/32

Once the interpretation is initialized (3.D.1), the first task is to get a coherent and
representative set of rate and pressure data. This includes loading the data (3.D.2), quality
check and validation (3.D.3), and editing to prepare for analysis (3.D.4). One or several
periods of interests, typically build-ups, will then be extracted and the diagnostic plot created
and the data diagnosed (3.D.5). The interpretation engineer can select one or several
candidate analytical and/or numerical models, set their parameters and generate these models
(3.D.6). For candidate models that are retained, the engineers can refine the parameters,
either manually or using nonlinear regression (3.D.7). Once the model parameters are
finalized, the user may assess the sensitivity and/or cross-correlations of the parameters using
confidence intervals from the nonlinear regression and run sensitivity analysis (3.D.8). Finally,
a report is issued (3.D.9).

The path above is the default path when all goes well. In reality, for complex problems, it may
be a trial-and-error process where the interpretation engineer may decide to go backwards
before continuing forward again when a segment of the process is not satisfactory.

3.D.1 Initialisation
The interpretation engineer must first input information required to identify the test and select
the main options that will set up the interpretation process: the type of fluid (that will
determine the function of the pressure to be used) and the type of test (production, DST,
interference, etc). The final input will be the parameters that are assumed to be known, and
that are required to calculate the interpretation results: porosity, net drained vertical
thickness, well radius, etc.

For a slightly compressible fluid, only a few PVT properties, assumed to be constant, are
needed: formation volume factor, viscosity, and total system compressibility.

For dry gas, a choice of correlations or input of PVT tables is required to calculate the
pseudopressure and pseudotime functions.

3.D.2 Loading Data


Import from flat ASCII files, manual input and copy-paste from a spreadsheet are the main
ways to load data. However, data is increasingly read from databases and/or with a direct real
time link to the running acquisition systems of service companies during the test.

Unlike for open-hole logs and despite several attempts in the past, as of today (2007) no
industry-standard ASCII format has emerged to execute the load with a click. In Canada the
EUB has published a format (PAS) for compulsory electronic submission of test data and
results, but this remains very oriented towards local legal procedures. So software packages
have to provide the engineer with the flexibility to interactively define the file format, in order
to cover the wide variety of existing files.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p17/32

Originally, the amount of data was a real issue because of limited available memory running
under DOS, the cost of computer memory, and the fact that the size of the data arrays had to
be declared and set by the software programmers at compilation time. All these limitations
have gone. Today’s software can easily handle multiple gauges and volume of gauge data
acquired during a well test which is rarely more than a million data points.

The recommended procedure is to load all data acquired from all gauges during the well test,
and not just a filtered subset. Filtering can always been applied later, on duplicate data sets.
However things are changing with the spread of permanent downhole gauges and the
increased usage of production data in Pressure Transient and Production Analysis. The amount
of production data is one order of magnitude higher, and is much less smooth than a typical
build-up. Smart filters, such as wavelets, are increasingly required to reduce the volume of the
data, retaining any trends and significant changes and eliminate noise. The processing of
Permanent Downhole Gauge (PDG) data is covered in Chapter 12.

3.D.3 Quality Control


Quality Control is an essential part of the interpretation, too often overlooked in the past, it
includes:

• Validation of the gauges: identification of failures, drift, clock failure, resolution, etc.

• Identification of operational problems.

• When applicable, identification and correction of tidal effects.

• Discrimination of wellbore effects from reservoir effects.

Quality Control has fortunately become a growing concern. Previously, an interpretation


engineer would often consider the build-up data only and match it with, for example, a radial
composite model; whilst all too readily ignoring the possibility that the test behavior could
indeed be related to wellbore phase redistribution. An efficient tool to diagnose wellbore effects
is the dynamic calculation of the difference between the gauges measuring the same data. The
value of the difference itself is of little interest as long as it stays within the gauges accuracy.
However, variations of the difference may be a valuable source of information.

When two pressure sensors are set at the same depth, as with a dual gauge carrier, their
difference can be used to check their synchronization (time shift) and their coherence. Gauge
failure and gauge drift may be identified.

When the gauges are set at different levels, as in tandem configuration, any change of the
pressure gradient occurring between the gauges may be detected. In the presence of phase
segregation problems, the proper placement of dual gauges may help qualifying and even
quantifying these problems. The engineer will avoid pointlessly interpreting and using a
complex reservoir model behavior that has absolutely nothing to do with the reservoir.

In the absence of dual gauges, one can calculate the derivative of the gauge versus time, and
plot it on a linear or log-log scale. This will act as a ‘magnifying glass’ of the pressure behavior.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p18/32

Fig. 3.D.1 - Drifting gauge and Data Quality control using dual gauges

Even though people associate the difficulty of well test interpretation to the modelling part, a
lot of thinking takes place at this stage of the interpretation, as it defines the starting point
from which the diagnostic will take place. Failing to identify any operational problems can
potentially jeopardize the whole interpretation process.

There is a side advantage in performing a comprehensive Quality Control: after going back and
forth between the data and the well test report, the interpretation engineer will know what
happened during the test, even if he/she was not on-site. See section 3.G for more details.

3.D.4 Editing data


Loaded data may be the result of careful post-processing by the data acquisition company, in
which case no or little editing may be needed. But very often the interpretation engineer will
have to gather data of unequal quality and from different sources. Pressures will often be
acquired downhole in real time or with a memory gauge, while rates will still be measured at
surface and come from the well test report with a different time sampling.

Beyond the usual cleaning of irrelevant data and the correction of load errors, the main
challenge will be to end up with at least one coherent, synchronized set of rate and pressure
data. To get there the engineer may have to perform the following tasks:

• Get all data acquired electronically to the same reference time.

• If not already loaded, create the rate history graphically by identifying the pressure breaks
and get the rate values from the well test report.

• Refine the production history, when the time sampling of rate measurements is too crude.

• Conversely, if the production history goes into useless details, simplify the rate history to
reduce the CPU time required to run the models.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p19/32

3.D.5 Extraction and diagnostic


Once the data have been synchronized and validated, the analysis itself will start. The engineer
will focus on one or several gauges, one or several flow periods, and will create the appropriate
diagnostic tools, starting with the loglog and the semilog plots. When several gauges are used,
they will be overlaid. When several production and/or shut-in periods are extracted, they will
be rate-normalized, then overlaid. In the case of Saphir, this extraction is followed by an
automatic positioning of a horizontal line for IARF on the Bourdet derivative and a unit slope
line for pure wellbore storage on both pressure and the Bourdet derivative. This positioning is
set by a relatively simple filtration, the main goal being to put these lines in the same ‘range’
as the data. Surprisingly, this sort of processing works quite well in the case of simple
responses, giving an instantaneous estimate of the wellbore storage and permeability-
thickness product. In case of complex behavior, the software may have selected the wrong
level of the derivative for IARF and or the wrong unit slope for wellbore storage. The
interpretation engineer will then interactively move the two straight lines in order to properly
position these flow regimes. See the illustration in Figure 3.D.2.

Fig. 3.D.2 – Match lines

During the extraction process, and possibly later, the engineer may decide to control the
derivative smoothing, apply a logarithmic filter, and in the case of a shut-in, control the last
flowing pressure.

After extraction, data problems overlooked in the initial quality control may become apparent,
requiring further data editing, and a new extraction.

Looking at the derivative response will generally be the starting point of this process.
Individual features in the derivative signature will be considered, validated or rejected, and
potentially associated to a well, reservoir or boundary model. These possible assumptions must
then be compared to what is already known from other sources.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p20/32

Depending on the diagnostic, the loglog and semilog plots can be complemented by other
specialized plots to identify specific flow regimes by straight line analysis. However this
approach has been made largely redundant by the introduction of the modern approach. The
engineer will have the choice of the pressure function, the time function and the type of
superposition that will be applied to the time function; raw function, tandem superposition for
simple build-ups, or multirate superposition.

Depending on the prior knowledge and the complexity of the response, the problem may be
very quickly restricted to one or two alternatives, or the range of possibilities may remain
large. For exploration wells, the uncertainty in the explanation may stand, and alternative
explanations may be presented in the ‘final’ report. Further tests and increased knowledge of
the reservoir could allow, later, narrowing down the range of possibilities, months or years
after the initial interpretation.

3.D.6 Model generation


The engineer, after diagnosing the behavior, will then select one or several candidate models.
The process below will be duplicated for each considered model.

The objective is to use the modelling capability of the software to match in part or in totality
the pressure response. This will consist of selecting one or several candidate models, that may
be analytical or numerical. Then entering a first estimate of the model parameters, running the
model and comparing the simulated results with the real data, on all relevant plots.

AI based Model advisers may be used to speed up the process by detecting if a derivative
response can be explained by a certain combination of well, reservoir and boundary models,
and produce a first estimate of the model parameters with no user interaction.

Today’s software products offer a wide choice of analytical models. Typically the user will
select a wellbore, a well, a reservoir and a boundary model. Unfortunately, our ability to solve
problems mathematically is limited, and all combinations of well, reservoir and boundary
models may not be available. This is sometimes frustrating to the engineer, as in this case only
portions of the response can be matched at any time.

There are many ways to estimate parameters: (1) from the results of specialized plots that
may have been created in the analysis; (2) from straight lines drawn on the loglog plot
(wellbore storage, IARF, fractures, closed systems, etc.); (3) from interactive features picking
the corresponding part of the derivative signature; (4) by manual input. For complex analytical
models, only a few parameters, or relationships between parameters, will be determined in a
unique way from the well test response. The other parameters or missing relations will be
input from other sources of information. If this missing information is not available, the
problem will remain under-specified.

The previous remark on parameter estimation is even more critical when using numerical
models, where the geometry will be essentially built from prior knowledge of the reservoir, and
only a few ‘global’ unknowns will be deduced from the test. Numerical models are presented in
further detail in Chapter 11.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p21/32

It is no longer a technical problem to transfer information and data directly and dynamically
between applications, and there are dozens of public or proprietary protocols to do so (OLE,
COM, DCOM, Corba, etc.). As a consequence models generated from third party applications
may be transferred and run in pressure transient analysis software. The most common
example is a ‘standard’ reservoir simulator run.

The model is generated and compared to the data, in terms of both pressure and Bourdet
derivative on the history plot, the loglog and semilog plots. In case other specialized plots are
used, the model will also be compared on these different scales. At this point, the engineer
may decide to reject the candidate model, or keep it and refine the parameter calculations.

3.D.7 Model refinement


Modification of the parameters: Before leaving the parameter refinement to an optimization
routine, the engineer should deal with the gross parameter errors if there are any. This will
increase the chance for the regression to succeed and converge faster, and it will secure the
choice of the model. Software will generally provide facilities to ease this process. For
example, parameters may be corrected if the engineer shifts the match on the loglog plot.
However, an experienced interpretation engineer with a good understanding of the sensitivity
to the model parameters will often do it faster by changing the values by hand.

Nonlinear regression: The principle is to use numerical optimization to refine the parameter
estimates by minimizing an error function, generally the standard deviation between the
simulated pressures and the real pressures at carefully selected times. The derivative may also
be integrated in the error function. The most commonly used optimization algorithm is
Levenberg-Marquardt, but there are many variants.

Among the model parameters, some may be fixed by the engineer. For the others, the
engineer may control their upper and lower limits. The data points on which the error function
will be calculated may also be user controlled. One major choice will be whether the
optimization is restricted to the analyzed period, or if it is extended to data outside the
analysis. In the first case, the match at the end of the optimization process will be as good as
or better than the starting point. If the optimization is performed on points beyond the
analyzed period, the overall history match will be improved, but potentially at the expense of
the quality of the match on the period used for the diagnostic.

At the end of the optimization process, the model is generated with the calculated parameters.
Figure 3.D.3 illustrates a final model match after optimization.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p22/32

Fig. 3.D.3 – Final match after optimization

3.D.8 Sensitivity study


At the end of the nonlinear regression it is also possible to recover the confidence intervals.
They can be used to assess the sensitivity to individual parameters and eventual parameters
cross-correlations.

Another possibility is to run and display series of model generations corresponding to different
values of a given parameter, in order to compare them on a single loglog plot. This is, to a
certain extent, the modern version of the type-curves, where dimensionless drawdown
solutions are replaced by the generation and extraction of detailed models with user preset
ranges of parameters. Figure 3.D.4 show the sensitivity to the distance between a well and
one single sealing fault.

Fig. 3.D.4 – Sensitivity to distance


Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p23/32

3.D.9 Reporting Guidelines


A typical interpretation report will be split into two components: the ‘mechanical’ part, basically
the result tables and plots generated, directly or indirectly, by the Pressure Transient Analysis
package, and the ‘verbose’ part, where the engineer will report the objectives, the operations,
the interpretation, the confidence one could have on his interpretation, and possible
recommendations for well treatments and/or future tests.

In 2005 there was no industry standard for reporting, except again the Canadian EUB
restricted to very basic results. Typically, professional interpretation reports will be generated
with two possible set-ups:

• A header document, from a word processor, with some ‘copy-paste’ of plots and results
from the PTA software, but with most of the ‘mechanical’ report delivered as an annex.

• An integrated document, typically from a word processor, where some plots and tables are
dynamically connected to the PTA software using some OLE or COM automations. The
advantage of this solution is that it is much more flexible. Once a model template has been
set up the reporting process will get shorter and shorter from one interpretation to the
next.

In any case the engineer must keep in mind that an interpretation is, at best, a guesstimate at
a given time, and truth can evolve with time. The key word here is ‘interpretation’.

‘The reservoir is a circle of radius 4123.93 ft’

This is probably the worse possible statement we can imagine in PTA. The reservoir is very
unlikely to be an exact circle. What we have in PTA is a range of models that we KNOW to be
over-simplified. We simplify to turn the qualitative into quantitative, and one must always be
factual. Also, the number of significant figures of a result must be reasonable, or at least not
ridiculous. It is not because the nonlinear regression finished at a given number that we must
keep all the significant figures of this number. So a much more reasonable statement would
be: ‘If we consider that the late time behavior is linked to a close system, a reasonable match
was obtained with a circular reservoir with a radius of 4,100 ft.

‘The more I know about well testing, the more I worry’. H.J. Ramey Jr, 1989

3.E Test design


Prior to any well testing, the objectives of the test must be defined, operational constraints
identified and scenarios produced that clearly show if the objectives will be met and if not, why
not.

In order to run the scenarios the engineer must rely on known information and data or make
reasonable assumptions. To explore ‘what-if’ scenarios sensitivity studies will be carried out
and based upon these the engineer will chose both downhole equipment including, amongst
others, pressure gauges along with their corresponding accuracy and resolution. In addition
surface equipment must be selected to include, amongst others, flowrate measurement
equipment. This done, the Engineer can make the safest and most economical plan to reach
the test objectives, this is a real test design.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p24/32

In order to illustrate the concept we will devise a scenario where the main objective of the test
is to determine skin, permeability and the distance to a seismic defined fault at some distance
away from the well (500 ft).

The analysis of nearby wells has indicated the nature of the reservoir to be homogenous with a
permeability-thickness of some 1000 mdft. The skin is of course unknown but believed to
become positive after drilling in this relatively high permeability rock. PVT parameters are
known in this saturated reservoir, and porosity and net drained thickness have easily been
determined from open hole logs in advance of the planned test.

It now remains for the engineer to design a program for the production of the well prior to the
build up planned for the analysis for kh, skin and boundary distance.

A simple, single flowrate was planned at 1000 stb/d for 24 hours followed by a build up of 100
hours, the 100 hours was chosen as the limit of build up time in this production well. The
reason is not necessarily to build up for the actual 100 hours. The fact is that this is a number
that, from most practical viewpoints is the limit of most build ups and allows the determination
of optimum build up time if the fault is indeed seen within this timeframe. Remember that the
diagnostic plot is in loglog coordinates. The test design option of the software was then used to
calculate the ‘what if’ sensitivity.

Because wellbore storage could also change, a ‘what if’ sensitivity on this parameter was also
performed. From Figure 3.E.1 it can be easily discerned that permeability can only be
determined with confidence if the skin is close to zero. That a wellbore storage coefficient
greater than 0.01 bbl/psi coupled with a skin of 10 has the same effect and any storage above
0.05 will mask the fault, see Figure 3.E.2. In case the assumption of zero skin holds and the
wellbore storage remains low then the build up time can be reduced to some 30 hours.

Fig. 3.E.1 – Skin sensitivity Fig. 3.E.2 – Wellbore storage sensitivity

Thus the engineer has to condition the well to reduce the skin and wellbore storage, maybe
through downhole shut-in.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p25/32

3.F Operational considerations: Gathering data


It is absolutely necessary, before running a test, to really identify what the test objectives are
and which data should be acquired. Part of this work will be achieved with the test design.
However, in complement the engineers and operators on site will need to clearly know what
information is needed.

Type of data Required?

Production history Yes

Pressure history Yes

PVT, correlations, tables or constraints Yes

rw Wellbore radius Yes

Φ Porosity Yes

h Net vertical drained thickness Yes

Field map with surrounding wells based on


Yes
seismic interpretation

Production history of surrounding wells Optional

Complete completion log preferably with a


Yes
permeability log. Core and core analysis

Completion diagram and geometry, deviation,


Yes
perfos, gauge depths.

All gauge, well test and operations reports Yes

Choice of flow correlations or availability of lift


Optional
curves from third party software
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p26/32

3.G Operational considerations: Validating data (QA/QC)


3.G.1 Introduction
During the early 1980’s the oil industry, especially in Northern Europe, went through a totally
justifiable drive to increase the quality and efficiency of work carried out in most of the
commonly related services involved in the production of oil and gas.

Quality control manuals and procedures had to be presented to prove that every effort was
fulfilled by the involved parties to assure that work, equipment and preparation procedures
were documented. These documents had to be approved by a ‘controlling body’ before the
company was qualified as an approved contractor. The QA/QC manuals described and
identified the procedures to follow, backed up by checklists, tests, inspections and certification
by approved inspectorates.

This drive was mainly directed towards offshore construction and sub-sea technology to
optimize equipment performance and the quality of engineering works, including the safety of
oil and gas installations and oil field operations.

The traditional service providers such as logging and well test companies were no exception,
but other than planned and regular calibration of mechanical and electronic measuring devices,
there was no attempt to set down documented criteria for measurement conditions and the
quality of the measured data.

As all service companies claim no responsibility for any loss or damage due to faulty
measurements and interpretations, procedures to validate measured data before any analysis
attempt was therefore not included or compulsory.

Today methods have been developed that allow quality control of downhole pressure and
temperature measurements to ensure that the data is valid and usable by the engineer for
further dynamic flow analysis (PTA and PA).

The validation increases the confidence in the results of the analysis and eliminates, to a large
degree, errors that could lead to major mistakes in the decision process of optimum
development and production of petroleum reserves.

This chapter addresses these procedures, and aims at teaching the techniques and assures
that measured data is valid and representative of the reservoir response.

3.G.2 Background
The introduction of the Bourdet derivative revolutionized the approach to pressure transient
analysis (PTA). This approach gave us a deeper visibility, multiplied our analysis ability but also
complicated the diagnostics by revealing phenomena unseen and not understood until this
time. The sensitivity to pressure changes, either caused by pure reservoir response or other
phenomena have become more accentuated, and the resulting interpretations more difficult to
justify since a variety of different analytical interpretation models could match the measured
data with widely different results.

Pressure behavior caused by fluid movements in the wellbore, phase segregation and
temperature anomalies, were often believed to be the result of a pure reservoir signal and
interpreted as such.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p27/32

Evidently, it was necessary to develop and employ a method to enable the engineer to
differentiate between the representative data of the reservoir response and the part caused as
a result of a signal from other phenomena.

The techniques described in this chapter were developed as a result of severe interpretation
problems encountered in many fields and led to the introduction of the concept of: ‘Differential
Pressure Analysis’.

It will be shown how the use of pressure differentials measured between pressure gauges
placed at different levels in the test string can help the interpreter to identify the pressure data
that is valid for interpretation, and enable him to save time by eliminating useless data caused
by anomalies and wellbore phenomena.

The method brings ‘Quality Control’ one step further, and should be an integral part in the
overall QA/QC programs of any company dealing with downhole measurements of pressure
and temperature.

3.G.3 The concept of ‘differential pressure analysis’


The analysis is based upon the difference in pressure measured between tandem pressure
gauges, the simplest case, or a combination of pressure differences if multiple gauges are used
during a pressure survey.

The study of these differences can reveal the following problems and has a direct impact on
the choice of the data measurements which are valid for pressure transient analysis (PTA):

• Phase segregation in the wellbore

• Fluid interface movements (oil, gas and water)

• Temperature anomalies affecting the pressure gauge and / or identification of gauges with
technical problems, such as:

o Pressure gauges outside of their claimed accuracy band and resolution specifications

o Gauge drift

o The response of the gauge to a depleting battery.

o Other technical or electronic malfunctions

By convention the pressure difference between gauges is calculated so that that an increase in
the ‘difference channel’ represents an increase in the fluid density between the gauges sensing
points, and a decrease a reduction of the fluid density, i.e.:

Δp = plower - pupper

The ‘difference channel’ behaviour will be identical whatever the gauge offset. The upper
gauge may well read a higher pressure than the lower gauge, possibly due to a gauge
problem, but the ‘difference channel’ would have the same identifiable shape.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p28/32

3.G.4 Basics
The simple analysis is based upon the study of the pressure and temperature differences
between two or more gauges placed in the test string at different depths. The figure below
shows schematically what happens to the pressure at the different sensors, if a ‘gas-oil’
interface is moving downwards and passes the sensors.

Fig. 3.G.1 - Fluid movement downwards

The example assumes that any ‘background’ behavior is following a constant transient or is in
‘pseudo steady state’ (PSS).

Once the ‘gas-oil’ interface hits the ‘upper sensor’ the pressure at this sensing point remains
‘constant’ as the interface moves towards the lower pressure point.

The pressure at the ‘lower sensor’ declines linearly if the fluid interface movement is constant
and becomes ‘constant’ again after the interface has moved below the lower pressure sensor.
The difference in pressure between the two sensing points is represented by the difference in
fluid gradient between oil and gas.

The following illustration represents the ‘difference channel’ between the two sensing points,
and by simple analysis it can be determined what fluid phase changes have caused the
observed phenomenon.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p29/32

Fig. 3.G.2 - Pressure difference

3.G.5 - Diagnosing phase segregation from differential pressure


The aim of the ‘Differential Pressure Analysis’ is to make an intelligent evaluation of any
observed anomaly, and attempt to identify approximately the different fluid phases in the
wellbore near the gauge sensing points. This will give a reasonable idea of wellbore
occurrences, and make it possible to calculate the fluid distribution from the primary gauge;
the gauge that has been identified as the most reliable and representative, and down to the
sandface. This is only an approximate calculation and cannot be used as a rule in all types of
wellbore phenomena, but it will nevertheless enhance the understanding and allow for an
intelligent correction of the gauge pressures from the gauge to a reservoir datum with more
confidence. This is a prerequisite for better confidence in material balance and identification of
drainage patterns.

Practically it involves a choice of several recognizable events on the difference channel (see
red circles below) and transfers the different delta pressure values to a simple spreadsheet.
Then estimations of fluid gradients can be made, based on already known fluid densities and
resulting ‘delta gradients’, that makes the total picture logical by applying a gauge offset.

Fig. 3.G.3 - QAQC plot with events marked


on difference channel
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p30/32

This method not only reveals wellbore anomalies but determines the pressure gauge offset,
which will have to be within the range of the gauge manufacturer’s claimed accuracy and
resolution.
DIFFERENTIAL PRESSURE ANALYSIS

Events From Difference Channel dp Assumed Assumed Calculated Implied Observed Residual
Time dp Gradient Fluid Gradient dp Offset dp difference
(hrs) (psi) (psi/ft) (psi/ft) (psi) calculated corrected (psi)
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10]

Event 1 17.10 6.30 Friction 0.714 6.510 -0.210 6.510 0.000


Event 2 17.18 2.90 -0.37 Oil 0.341 3.110 -0.210 3.110 0.000
Event 3 17.33 0.20 #REF! Gas 0.045 0.410 -0.210 0.410 0.000

Distance between sensors: 9.12 ft Assumed Accuracy Offset: -0.21 psi

Column (2) and (3) are read directly from the difference channel from the quality control plot.
Column (4) is the differential gradient calculated from column (3). Column (5) and (6) are
intelligent guesses or assumptions by the user that will normalize column (8), i.e. the same
implied gauge offset applies whatever the assumed fluid phase is present.

The implied offset is then entered by the user in its appropriate box below the table, and the
residual differences will then become close to zero, or zero, if the correct assumption as to the
fluids present in the wellbore have been made.

The implied offset becomes the gauge offset which has to be within the gauge specifications to
be acceptable to the operator.

This analysis gives a much better idea as to which of the fluid gradients to use to correct the
gauge pressure to top of sandface.

Another product of this approach is being able to decide which part of the data is really valid
for pressure transient analysis as all data affected by the segregation or fluid movements must
be discounted in the interpretation process. The impact of not knowing and understanding the
wellbore environment is now described.

3.G.6 Impact on diagnostics


The impact on the diagnostics and the choice of model is illustrated below. The upper gauge is
affected by phase segregation and the shape of the derivative is clearly distorted resulting in a
wrong choice of the interpretation model which is straight homogeneous as illustrated by the
response of the lower gauge.
Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p31/32

Fig. 3.G.4 - Lower gauge; homogeneous Fig. 3.G.5 - Upper gauge; double porosity

3.G.7 Gauge drift


Gauge drift is caused by unstable electronic components and fatigue of the sensing material
used in the instruments. Strain gauges are particular susceptible to this problem.

Drift during a relatively short well test is uncommon, especially today as the quality of
electronic gauges have increased immensely compared to those available 10-15 years ago.
However, it does happen, and if the drift is severe, can easily lead to wrong diagnostics during
PTA.

The drift problem is more common during long measurements and can therefore be a real
problem in permanent downhole gauges (PDG).

In both cases (short well tests or PDG) it is important to check the data for validity through the
QA/QC procedures described in this document before attempting serious analysis. To identify a
drifting gauge it is necessary to have two or more measurements and study the difference
channel constructed between a reference gauge and all the other gauges.

The figures below illustrate well the difference channel and the impact on the analysis of the
drifting gauge without the prior knowledge that such drift exists.

Fig. 3.G.6 – QA/QC plot with difference


Dynamic Flow Analysis - v4.02 - © KAPPA 1988-2007 Chapter 3 – Pressure Transient Analysis - p32/32

Fig. 3.G.7 - Non drifting gauge, no fault Fig. 3.G.8 - Drifting gauge, with fault

You might also like