Vous êtes sur la page 1sur 22

Schlumberger

D.1 PRODUCTION EVALUATION


D.1.1

INTRODUCTION

Production logging measurements provide


the operator with detailed information on the
nature and behaviour of fluids in the well
during production or injection. Major applications of production logging include:
evaluating completion efficiency
detecting mechanical problems, breakthrough, coning
providing guidance for workovers, enhanced recovery projects
evaluating treatment effectiveness
monitoring and profiling of production
and injection
detecting thief zones, channelled cement
single layer and multiple layer well test
evaluation
determining reservoir characteristics
identifying reservoir boundaries for
field development
There is a basic family of production logging tools, designed specifically for measuring the performance of producing and injection wells. The sensors included are:
temperature
fluid density (gradiomanometer, nuclear)
holdup meter
flowmeter spinners (continuous, fullbore, diverter)
Manometer (strain gauge, quartz gauge)
caliper
noise (single frequency, multiple frequency)
radioactive tracer

Many of these sensors can be combined into


one tool and recorded simultaneously to
measure fluid entries and exits, standing liquid levels, bottomhole flowing and shut-in
pressures, pressure losses in the tubing, and
the integrity of the gravel pack and hardware
assemblies. Since the measurements are
made simultaneously, their correlation is less
affected by any well instability that might
cause downhole conditions to vary over a period of time. The tool string also includes a
casing collar locator and a gamma ray tool for
correlation and depth control. Figure D1
shows a schematic of the sensors in a typical
production logging tool string.

Fig. D1: PLT, simultaneous production logging tool.

(11/96) D-1

Introduction to Cased Hole Logging

Production Logging Applications


A great value of production logs lies in their
ability to provide determinations of the dynamic flow patterns of well fluids under stable producing or injecting conditions. For a
number of reasons production data from other
sources may be misleading. Some of these
reasons are:
Surface measurements of pressures,
temperatures, and flow rates are not
necessarily diagnostic of what is happening in the well.
Fluid flow outside the presumed paths,
such as through cement channels in the
annulus, can only be detected by production logs.
Zone-by-zone measurements of perforating efficiency, impractical except by
production logs, are often necessary to
identify the actual producing or receiving intervals.
Zone-by-zone measurements of pressure and flow rate can be used to determine the average pressure and the
productivity index of each producing or
injected interval.
Production logs therefore have useful application in two broad areas: evaluation of well
performance with respect to the reservoir and
analysis of mechanical problems.
Well Performance
In a producing well, production logs can
determine which perforated zones are giving
up fluids, ascertain the types and proportions
of the fluids, and measure the downhole conditions of temperature and pressure, and the
rates at which the fluids are flowing. If thief
zones or other unwanted downhole fluid circulations exist, they can be pinpointed.

(11/96) D-2

Injection wells are especially well adapted


to production log analysis because the fluid is
monophasic and of a known and controlled
type. The objective of logging is to locate the
zones taking fluid and to detect lost injection
through the casing annulus.
Job Planning
Planning is the most important facet of a
successful production logging job. Close coordination between Schlumberger engineers
and well operators is essential.
Planning should start with defining and
analyzing the expected downhole injection or
production rates, pressures, temperatures, and
fluid types. This analysis will determine the
tool types and resolutions needed to solve the
problem. The presence of H2S and CO2
should also be considered. The following information is required to plan the operation:
a detailed well sketch
Christmas tree specifications for rigup
sand or formation fines production
presence of paraffin or scale deposits
knowledge of whether the well was hydraulically fractured and/or acidized
frac balls usage
reservoir data, reservoir and fluid
properties
production history.
Before the production logging operation is
attempted, the operator should verify that the
well conditions are acceptable by running a
dummy tool to the bottom to determine if
there are any obstacles. Any problems should
be remedied before the logging operation is
started.

Schlumberger

Time allocation is an important consideration for production logging operations - particularly in high pressure operations. Surveys
can frequently be run more safely in daylight.
If not this may dictate the use of special
lighting equipment for lengthy operations.
Well Problems
In the absence of knowledge to the contrary,
it is assumed that the well has hydraulic integrity, and that the fluids are going where they
belong; often, this assumption is wrong. Examples include: casing leaks, tubing leaks,
packer leaks, communication through the annulus due to poor cement, and thief zones.
Figure D2 shows how these conditions can
lead to misleading conclusions when well performance data come from surface measurements alone. Solutions to these and other
well problems can be found by the integration
and interpretation of production log data.
Fig. D2: Mechanical well problems.

We will cover the various production logging tools and how each is interpreted individually. Once the measurement made by
each PL sensor is understood we are then
better able to combine them for a better interpretation. In the interpretation section we will
cover the various flow regimes occurring in
the wellbore during production.

(11/96) D-3

Introduction to Cased Hole Logging

(11/96) D-4

Schlumberger

D.2

FLOWMETER TOOLS

Flow Velocity
Flowmeter spinner tools and radioactive
tracer tools are usually used to measure flow
velocity. Under certain conditions, the fluid
density and temperature tools can be used to
estimate flow rates but their use for this purpose is much less common.
Spinner Flowmeter Tools
Spinner flowmeters all incorporate some
type of impeller that is rotated by fluid moving relative to the impeller. The impeller
commonly turns on a shaft with magnets that
rotate inside a coil. The induced current in
the coil is monitored and converted to a spinner speed in revolutions per second. This
spinner speed is then converted to fluid velocity (flow rate).
D.2.1

CONTINUOUS FLOWMETER
TOOL

The continuous flowmeter tool has an impeller mounted inside the tool, or in some versions, at the end of it (Figure D3). The most
common tool diameter is 42.9 mm (1 11/16 in)
with the spinner being smaller. The continuous flowmeter is most often run in tubing
where the fluid velocities are high and the
fluids tend to be a homogeneous mixture.
The spinner covers a much larger percentage
of the cross-sectional flowing area than in
casing and tends to average the fluid velocity
profile.

Fig.D3: Continuous Flowmeter.

(11/96) D-5

Introduction to Cased Hole Logging

Fig. D5: Automatic Diverter Flowmeter tool.

Fig.D4: Fullbore spinner flowmeter tool.

D.2.2

FULLBORE SPINNER TOOL


(FBS)

The FBS tool is probably the most commonly run spinner tool. The tool collapses to
traverse the tubing and opens inside casing for
logging purposes. The large cross-sectional
area of the spinner tends to correct for fluid
velocity profiles and multiphase flow effects.
A schematic of the FBS tool, in both the collapsed, through-tubing and opened, belowtubing, configuration is shown in Figure D4.

(11/96) D-6

D.2.3

AUTOMATIC DIVERTER
FLOWMETER TOOL (ADF)

The automatic diverter flowmeter tool utilizes a fabric diverter with an inflatable ring
for use in medium and low flow rate wells.
This diverter assembly fits inside a metal cage
that is closed and protects the diverter while
entering the well. The metal cage is opened
and closed on command from the surface and,
when open, helps to centralize the tool and
deploy the diverter. At the same time, fluid
carried with the tool is pumped into the inflatable ring, thus obtaining a seal to the casing.
A schematic of the automatic diverter flowmeter tool is shown in Figure D5.

Schlumberger

Fig D6: Typical production logging


tool string used for testing.

The automatic diverter flowmeter tool has


good fluid velocity characteristics since all of
the fluids moving through the casing must
pass through the spinner section. It is particularly appropriate for multiphase flow since
the fullbore spinner measurement can be adversely affected by the downflow of the
heavier phase.
The tool can be combined with other production logging sensors (Figure D6) so that
both a continuous flow profile and accurate
station measurements can be made on the
same survey. This is particularly useful for
logging in layered reservoirs with multiple
perf zones.

Fig D7: Typical fullbore flowmeter log.

D.2.4

INTERPRETATION OF
SINGLE PHASE FLOW USING
SPINNERS
Figure D7 shows a typical Fullbore Flowmeter Log. It displays CVEL (cable velocity)
in Track #1 (left side) and spinner response
RPS (revolution per second) in Tracks #2 and
#3.

(11/96) D-7

Introduction to Cased Hole Logging

Fig D8: Determining flow rates


in single-phase flow.

The average lines should be approximately


parallel to the static fluid column lines. A
shut-in run is helpful to obtain the slope of the
response lines if a static column does not exist.
Fig D9: Spinner calibration plot.

For Percentage Contribution of Each Zone


Spinner revolution rate varies with fluid
flow rate and the relationship is generally linear for continuous flowmeters, the fullbore
spinner tool, and the fluid diverter tool.
Therefore, in single-phase flow, the flow profiling interpretation technique is essentially
the plotting of spinner data in revolutions per
second, such that the percentage flow contribution of each zone can be read directly from
the plot. This assumes the fluid density and
viscosity are consistent throughout the interval and that the velocity profile does not
change. An example showing percentage
contributions is shown in Figure D8.

(11/96) D-8

For Absolute Flow Rates


Spinner rate is a function of fluid viscosity,
density, and velocity. Care must be taken if
absolute flow rates, rather than percentage
contributions, are desired from the flowmeter
data or if percentage contributions are desired
in an interval with varying viscosity or density. Under these conditions, downhole calibrations for continuous data are used for determining absolute flow rates. This is true
even for single-phase flow.
When performing downhole calibrations for
absolute flowrates, a plot is constructed using
several logging passes. The passes should be
both up and down at various cable speeds.
The cable speed in meters per minute is plotted on the x-axis for various passes (as in Figure D9). The down passes are positive meters
per minute; the up passes are negative.

Schlumberger

The associated spinner speeds are plotted on


the y-axis. The down passes in producing
wells are positive revolutions per second
(RPS). The up passes are negative, if logged
faster than fluid flow, and positive, if logged
slower than fluid flow.
Typically, a plot is constructed from data
above a set of perforations (stabilized flow) or
in the sump (area of no flow). The plot above
shows two lines drawn on the RPS versus VT
graph. In a static column (area of no flow) the
data will construct a line that comes close to
crossing through the origin (dashed line). The
line does not cross directly through the origin
because of nonlinearity at low rates and the
threshold velocity required to initiate the propeller turning. This single line then becomes a
calibration chart that can be entered with RPS
from the log at any point. The meters per minute value associated with an RPS value on
the plot is the total velocity turning the spinner. After subtracting the cable speed for the
pass being interpreted, the net velocity represents the fluid velocity portion of the spinner
speed. This value must be multiplied by the
velocity profile factor, which is usually taken
to be 0.83, to represent the actual average
fluid velocity in the casing. The flowrate can
be determined by consulting the appropriate
chart for the relationship between fluid flow
volumes and fluid velocity for a specific casing inside diameter (Table D1).
A second line can be drawn on the same
plot from data above a set of perforations
(area of stabilized flow). This data will plot to
the northwest of the static column line for
flow up the well (solid line on the plot). The
fluid velocity can be found directly from the
distance between the static column intercept
with RPS = 0 (typically the origin) and the
flowing line at RPS = 0. In this case the fluid

velocity can be read directly from the VT scale.


Tool velocity will be compensated by using
the intercept method.
Note that for contributing zones the fluid
velocity will be positive (VT Up scale). For
thief zones, the data will plot to the southeast
of the static column and fluid velocity will be
read on the VT Down scale or negative scale.

Fig. D10: Two-pass interpretation technique.

Two-Pass Technique
The two-pass technique can be used to calculate the percent contribution of each zone in
varying viscosity conditions, whether from
multiphase flow or single-phase flow with
multiple viscosities. This technique consists
of running several continuous flowmeter
passes against the flow direction and with the
flow direction. The cable speed must be
faster than the fluid velocity on the passes
with the fluid flow direction. Two passes, one
with and one against the flow, are selected
and then normalized to coincide in a region of
no-flow (i.e., below all perforations). The
amount of separation measured in log divisions between the two passes after normalizing is linearly proportional to fluid velocity.
One hundred percent flow is at the point of

(11/96) D-9

Introduction to Cased Hole Logging

maximum deflection, which is usually above


all perforations. Thief zones complicate the
interpretation somewhat, but the principle remains the same.
A distinct advantage of this technique is that
it cancels the effect of viscosity changes.
These changes are essentially shifts in RPS
readings in the same amount and direction on
both passes. Thus, the separation remains independent of viscosity effects. If the centerline is defined as a line halfway between the
two curves, a centerline shift to the right is a
viscosity decrease; a centerline shift to the left
is a viscosity increase (Figure D10). If absolute fluid velocity is desired from the two-pass
technique, and if multiple calibration passes
have been run, it can be computed from the
following equation:
vf = 0.83 [(RPS) / (Bu + Bd)],
where:
Bu is the up calibration line slope in
rps per meter per minute.
Bd is the down calibration line slope
in rps per meter per minute.
Bu and Bd can, and often will, be
slightly different numerically.
RPS is the amount of separation
between up and down passes in RPS.
Although the foregoing comments focus on
fluid
viscosity
changes,
the
effects/assumptions regarding fluid density
changes are similar but opposite in effect.
Fluid velocity can be converted to flow rates
in barrels per day with production log chart 610 (see Table D1).

(11/96) D-10

Fig. D11: Two-pass overlay technique.


(Composite Spinner log)

An actual two-pass overlay technique is


shown in Figure D11. Things to note are:
1. Difference between the up and down
overlays. The spinner is reversing on
three of the up passes.
2. Curves overlay in the sump where total
flow is zero.
3. Curves overlay above the top perforations where the flow is maximum.
4. The flowrate varies with the separation
between spinner passes. i.e. the maximum contribution occurs where the
curves diverge the most (2940 to
2946m).

Schlumberger

D.3 TRACER TOOLS


Tracer tools can be placed into two basic
categories. These are:
1. Gamma ray tools that do not have downhole ejectors for releasing radioactive material, and
2. Gamma ray tools that have downhole
ejectors in combination with multiple
gamma ray detectors.
The first category is comprised of tools that
are essentially the same as those used for
openhole logging. These are usually smaller
diameter tools for through-tubing application.
The more common sizes are 35mm and
43mm. In addition to flow profiling with the
controlled time technique and traditional
openhole logging, these tools are often used
for channel detection by comparing logging
runs made before and after injecting fluids
containing radioactive material into the well.
The difference in the two runs will identify
where radioactive materials are present. If
radioactive material is present at any point
other than the perforated intervals, channeling
or vertical fracturing is likely.
Ejecting tools generally consist of two basic
downhole components. The first component
is a chamber that will hold a small amount of
radioactive fluid and will eject a controlled
amount of this material into the borehole.
The second component is a multiple detector
system that can monitor the movement and
location of the tracer fluid that has been released. The types of ejectors and detector
systems vary with tool application and sophistication. A popular tool configuration
consists of one ejector and three detectors (see
Figure D12).

Fig. D12: Typical tracer tool configuration utilizing


one ejector plus three detectors.

(11/96) D-11

Introduction to Cased Hole Logging

The tool configuration depends on the fluid


flow direction. If logging an injection well,
the configuration will normally be one detector above the ejector and two spaced detectors
below. In a producing situation, two detectors
are placed above the ejector and one detector
is placed below. The purpose of the single
detector on the opposite side of the ejector
from the flow direction is for detecting unexpected flow reversals produced by thief zones
and for identifying channels behind casing,
where fluid is flowing opposite the direction
of the wellbore fluids. The purpose of the two
adjacent detectors is for flow profiling as a
function of flow time between the two detectors.

Tracer Flow Rate Formula:


spacing x (/4)(dh2-dtool2) x 1440
q=
time x (1/60)
Where:
q is the flowrate in m3/day.
spacing is the distance between
the two detectors in meters.
dh is the casing inside diameter
in meters.
dtool is the tool diameter in meters.
time is the elapsed time between
detections in seconds.
For example:

The principle of ejector tracer logging is the


releasing of a radioactive isotope that dissolves in the wellbore and becomes part of the
wellbore fluid. The tracer material moves at
the same velocity as the wellbore fluid. A
measurement of the elapsed detection time
between the two detectors, along with knowledge of the tool configuration, is enough information for computing fluid flowrate. This
assumes, of course, that the tool is not moving. Unlike the controlled time survey, the
tool diameter must be considered in the
flowrate computation because it subtracts
from the casing internal cross-sectional area.

(11/96) D-12

Casing ID = 161.700mm = 0.161700m.


Tool OD = 34.925mm = 0.034925m.
For near detector to far detector, with spacing 1.4986m, and elapsed detection time of 15
seconds:
q = 169 m3/day
The sensitivity of the detectors to gamma
rays allows the system to monitor radiation
changes inside the casing wall and outside the
casing near the casing wall. The actual depth
of investigation of the gamma ray detector
depends on the type of detector, and the magnitude of the radiation. In most cases, it can
be estimated at 0.3 meters.

Schlumberger

Water, oil, or gas soluble tracer materials


can be used. Water soluble material is the
most common.
Some fluid flow applications of radioactive
tracer logging are:
- To check for packer, casing, or tubing
leaks;
- To identify channeling;
- To establish injection profiles on injector wells;
- To imply production profiles from injection profiles on production wells during
injection testing; and
- To establish flow profiles in low flow areas
of producing wells. (Tracer logging in
producing wells requires special considerations.)
Fig.D13: Single shot of a RA Tracer Tool.

Most of these applications require logging


techniques and interpretation methods unique
to the problem.
Radioactive tracer surveys are not routinely
run in producing wells because of the complications of produced radioactive fluid and
multiphase flow effects. Therefore, the main
application of this technique is in injection
wells.
As a general rule, the flowmeter gives more
accurate results in high flow rates and the radioactive tracer technique provides better results in flow rates less than about 100 B/D.

5.3.1 TRACER LOG INTERPRETATION USING DATA FROM


EJECTOR TOOLS
(Velocity Shots)
Figure D13 is a velocity shot of a Radioactive Tracer. The tool is stationary in the hole
and a radioactive slug is ejected from the tool.
A gamma ray detector is above the ejector and
two below. Thus, as we can see from our example, when a slug is ejected into the wellbore, it moves past the nearest lower detector
(GRTE), which is the solid trace in Tracks II
and III. The slug then passes the lowest detector (GRSG), which is the dashed trace in
Tracks II and III.

(11/96) D-13

Introduction to Cased Hole Logging

1. Into the formation, sand #3, #2


2. Into an "up" channel behind pipe, from
sand #3 into sand #4.
3. Into a "down" channel behind pipe,
from sand #2 into sand #1.
If the tool is immediately above the perforations, a slug entering an upward channel will
be seen by all three detectors, depending on
where the slug enters the formation. The slug,
if seen, will be of much lesser amplitude and
will be poorly defined since only a fraction of
the tracer material may enter the channel. The
slug may only cause a gentle deflection of one
division for maybe six seconds. To check for
a channel below perforations, the lowermost
detector is placed below the perforations. The
next lowest detector is placed in the perforations. Thus, if a channel exists, the slug will
be seen on the lowermost detector. Storage of
gamma ray material may also be indicative of
channelling. This can be seen by ejecting a
large slug and then logging the borehole to
determine which zone is taking the fluid.
Fig. D14: Potential fluid paths.

The slug then enters the perforations. Since


we know the time, usually two seconds per
each horizontal "depth" line, 10 seconds for
bold horizontal lines, the time and distance
the slug passes between the detectors is
known and thus fluid velocity is obtained. As
the slug passes into the perforations, it can
follow three paths (Figure D14):

(11/96) D-14

When the Radioactive Tracer is run in combination with a temperature log they can be
combined as in Figure D15. By presenting
injecting temperatures, storage temperatures
and radioactivity passes the presence of a
channel can be confirmed. Relative amounts
of storage are dependent on permeability and
hence should not be interpreted from a volume standpoint.

Schlumberger

Controlled Time Survey


The controlled time method qualitatively
detects the flow of fluids up or down the hole,
either in the casing or in the annulus. Figure 5
shows an example of the controlled-time Radioactive Tracer Survey. In this case radioactive material was ejected at the bottom of the
tubing and successive runs were made with
the gamma ray tool. The times of the injection and of each log run were carefully noted.

The radioactive slug (points a, c, e, and h)


may be seen to move down the casing. After
entering the perforations opposite sand 3, a
part of the radioactive slug (points f, j, h, and
v) channels up the casing annulus to sand 4.
After entering at sand 2, part of the radioactive slug (points l and p) channels down the
casing annulus to sand 1. Fluid appears to be
entering sand 3 because of the stationary
readings at points i, m, and q. And finally,
some radioactive material is trapped in a turbulence pattern just below the tubing as
shown by points b, d, g, and k.

(11/96) D-15

Introduction to Cased Hole Logging

Fig. D15: Combination Temperature and Radioactive Tracer Log.

(11/96) D-16

Schlumberger

Fig. D16: Radioactive tracer survey: timed runs analysis.

(11/96) D-17

Introduction to Cased Hole Logging

(11/96) D-18

Schlumberger

D.4 FLUID DENSITY TOOLS


There are two major types of fluid density
tools:
1. Gradiomanometer fluid density tool
2. Nuclear fluid density tool (gamma ray absorption)
A third tool type works on a principle other
than fluid density, it is the capacitance or watercut tool.
In a two-phase system (e.g., gas and water),
knowledge of the downhole density of each
phase, plus the mixture density from the fluid
density survey gives the log analyst the percent (holdup) of each phase occupying the
casing at the point of interest. Once the
holdup is determined from the fluid density
survey, it can be used to find the flowrate of
each phase, assuming the mixture flowrate is
known from a velocity survey (spinner) and
the slippage velocity is known. Slippage velocity is the difference between the rise rate of
the two fluids due to the difference in their
downhole densities. The following case demonstrates the basic interpretation technique for
two-phase flow. Use of watercut information
from a capacitance tool is very similar to use
of fluid density readings for interpreting twophase flow.

Fig. D17: (Chart 6-15) Liquid Slippage Velocities.

For slippage velocity determination, use


Chart 6-15 for liquids. In gas wells use 60
ft/min, if no other information is available.

(11/96) D-19

Introduction to Cased Hole Logging

D.4.1 THE GRADIOMANOMETER


FLUID DENSITY TOOL
The Gradiomanometer tool uses the pressure differential between two pressure sensors
spaced a known distance apart; e.g., two feet;
to infer the density of the fluid between the
sensors. There are several types of pressure
sensors that can be used in the gradiomanometer application.

Fig. D18: Gradiomanometer fluid density tool.

The example tool shown in Figure D18 uses


a bellows system. The bellows will compress
with pressure. The lower set of bellows will
be slightly more compressed than the upper
set. The mechanical linkage between the
bellows is constructed such that a rod moves
in proportion to the difference in compression
between the two sets of bellows. A magnetic
plunger on the end of the rod generates a
signal in the transducer coil proportional to
the rod movement. This allows the coil output
to be calibrated in terms of fluid density.

HOLDUP = Y
YHP - Holdup of the heavy phase
YLP - Holdup of the light phase
log -Tool Response at any point in the well.
LP - Downhole density of light phase.
HP - Downhole density of heavy phase.
log - LP
YHP =

YLP =

HP - LP
1 - YHP

HP and LP are obtained from shut in fluid


density passes if possible.

(11/96) D-20

Figure D19 shows a Composite Gradiomanometer log. The initial shut-in pass indicates
water in the sump (TD to 2948); an oil leg
(2948 to 2920); finally gas (from 2920 up).
The flowing pass shows the entry points for
each fluid.

Schlumberger

compared to the nuclear systems that exhibit


statistical variations.
A disadvantage of the pressure differential
system is the fluid flow around the tool can
cause friction effects that alter the apparent
pressure differential, which produces erroneous fluid density readings.
D.4.2

THE NUCLEAR DENSITY


TOOL

The nuclear fluid density tool operates on a


similar principle to the formation density
tools; i.e., a source of gamma rays is positioned with respect to a detector of gamma
rays so that the wellbore fluid acts as an absorber. Figure D20 illustrates this operating
principle. A high count rate indicates a fluid
of low density, and a low count rate indicates
a fluid of high density.

Fig. D19: Typical Composite Gradiomanometer log.

When a well is deviated, the gradiomanometer reading must be divided by the cosine of the deviation angle to correct for the
hole deviation effect.
An advantage of the pressure differential
system is that it has a very smooth readout

The advantage of the nuclear fluid density


tool over the gradiomanometer is that its
measurement is not affected by wellbore deviation or by friction effects. However, since
the tool relies on radioactive decay, the readings are subject to statistical variations. It
should also be noted that the measured quantity is the average density of the flowing
mixture; thus, it is subject to the same holdup
effects as the gradiomanometer.

Fig. D20: Nuclear fluid density tool.

(11/96) D-21

Introduction to Cased Hole Logging

Fig. D21: Capacitance tool.

D.4.3

THE CAPACITANCE
(DIELECTRIC OR WATERCUT) TOOL

The third group of widely used tools for


distinguishing water from hydrocarbons depend for their operation on the difference
between the dielectric constant of water (80)
and that of oil or gas (6). A simple way to
find the dielectric constant of a fluid is to use
the fluid as the dielectric between the plates of
a capacitor. The capacitance may be found by
classical methods such as including it in an
RC network and finding the resonant frequency.

(11/96) D-22

A conventional design is shown in Figure


D21. Two cylindrical metal tubes are arranged so that wellbore fluids flow through
the annular space between them. The raw
readings of such a device are in terms of a
frequency. Each tool will have a calibration
graph to convert a measured frequency to a
watercut value. These tools behave well, provided that the continuous phase is oil. In
practice, the measurement may become unreliable if the watercut in the flowing mixture
exceeds 30%.

Vous aimerez peut-être aussi