WellPlacementFundamentals Aug2008 1
WellPlacementFundamentals Aug2008 1
Roger Griffiths
1 Overview.................................................................................................................................. 5
1.1 Definition ........................................................................................................................... 5
1.2 Introduction........................................................................................................................ 5
1.3 Well placement economics ............................................................................................... 6
1.4 Well location terminology .................................................................................................. 7
1.5 The three complementary methods of well placement ................................................... 10
1.5.1 Model, compare, and update ................................................................................... 12
1.5.1.1 Building the formation model ............................................................................ 12
1.5.1.2 Computing tool responses................................................................................ 13
1.5.1.3 Real-time comparing and model updating........................................................ 16
1.5.2 Real-time dip determination from azimuthal measurements ................................... 21
1.5.2.1 Limitations of nonazimuthal measurements ..................................................... 21
1.5.2.2 Azimuthal measurements ................................................................................. 22
1.5.2.3 Image color scaling........................................................................................... 27
1.5.2.4 Image acquisition considerations ..................................................................... 27
1.5.2.5 Using images to calculate the incidence angle between the borehole and a
layer 30
1.5.2.6 Strike, true dip, and azimuth............................................................................. 31
1.5.2.7 Apparent dip and incidence angle .................................................................... 31
1.5.2.8 Image orientation for formation dip determination............................................ 33
1.5.2.9 3D visualization................................................................................................. 33
1.5.2.10 Using quadrant data to calculate incidence angles .......................................... 34
1.5.2.11 Using dip to calculate layer thickness............................................................... 36
1.5.2.12 Projecting to the bit ........................................................................................... 38
1.5.3 Remote detection of boundaries.............................................................................. 39
1.6 The three components of well placement ....................................................................... 41
1.6.1 Downhole tools ........................................................................................................ 41
1.6.1.1 Directional drilling technology........................................................................... 41
1.6.1.2 Measurement-while-drilling technology ............................................................ 47
1.6.1.3 Logging-while-drilling technology ..................................................................... 49
1.6.2 Software and information technology ...................................................................... 51
1.6.2.1 Real-time data transmission ............................................................................. 51
1.6.2.2 Real-time information extraction....................................................................... 53
1.6.3 People and process ................................................................................................. 55
1.6.3.1 The role of operating company geoscientists................................................... 55
1.6.3.2 The role of the well placement engineer .......................................................... 56
1.6.3.3 Team coordination ............................................................................................ 56
2 Reservoir Geology Fundamentals ...................................................................................... 57
2.1 Rock classifications......................................................................................................... 57
2.1.1 Igneous .................................................................................................................... 57
2.1.2 Metamorphic ............................................................................................................ 57
2.1.3 Sedimentary............................................................................................................. 57
2.2 The rock cycle ................................................................................................................. 58
2.3 Depositional environments.............................................................................................. 59
2.4 Plate tectonics ................................................................................................................. 60
2.4.1 Divergent margins.................................................................................................... 61
2.4.2 Convergent margins................................................................................................. 62
2.4.3 Strike-slip margins ................................................................................................... 62
2.5 Faults............................................................................................................................... 62
The well placement process is an interactive approach to well construction, combining technology
and people to optimally place wellbores in a given geological setting to maximize production or
injection performance. Accurate well placement helps improve the return on the money invested
in drilling the well.
1.2 Introduction
The success of a well can be measured both in the short and long term:
• In the short term the success of the well is determined by whether it is drilled safely,
efficiently, on time, and on budget and is producing hydrocarbons at the expected rate or
better.
• In the long term considerations such as access to reserves, delayed onset of water
production, extended production, and reduced intervention costs determine the total revenue
generated from the well and hence the return on investment from drilling the well.
Well placement improves both the long-term and short-term performance of a well (Fig. 1-1). The
drilling rate of penetration (ROP) is generally improved because the well remains in the more
porous reservoir, which can be drilled faster than the surrounding formation, and sidetracks are
avoided. By staying in the reservoir rather than the nonproductive surrounding formations,
production is also improved.
The key to maximizing reserves recovery is placement of the well in the reservoir such that it
produces hydrocarbons for the longest possible time and drains the formation as completely as
possible. By accessing pockets of untapped formation and avoiding unwanted fluids, a well
ultimately delivers the maximum return on investment by delivering the maximum possible
hydrocarbon volume with the minimum associated water or unwanted gas production.
Cumulative
Hydrocarbon
Production Accurately placed well
Increased
Increased recovery
production
rate
Conventional well
Time
Improved well construction efficiency Sustained high production rate
Earlier production
Figure 1-1. Well placement improves well and asset economics by improving well
construction efficiency, delivering higher sustained hydrocarbon production rates and
improving hydrocarbon recovery.
2-3 m
Sweet Capture Attic
Attic oilO il
Spot
Figure 1-2. Optimal well placement (blue solid line) ensures short-term production, long-
term reserves access, and delayed onset of water production. The red dashed line shows
an undesirable, unintended exit into the overlying shale. Drilling too low in the formation
(black dashed line) is also undesirable because it leaves attic oil when the formation is
drained.
Horizontal wells generally have higher risk than vertical wells. The requirement to keep the well in
the productive interval demands improved directional drilling and reservoir understanding beyond
that required to drill vertically into a target. Where these improvements are not in place prior to
drilling horizontally the increased expenditure may not deliver improved well performance
because the target zone is not exposed by the well as assumed in the well prognosis.
The well placement techniques outlined in the following chapters help mitigate the risks
associated with positioning the well in the target. By improving drilling efficiency, reservoir
contact, productivity, and profitability, well placement can deliver the horizontal well potential that
has eluded some operators.
Return
$ well #2
0
well #1
Cost
Time
Figure 1-3. Return on investment in drilling a well is dependent on both the rate and
duration of production.
Accurate well placement helps improve well construction efficiency, reduce drilling risk, extend
reservoir contact, maximize reservoir exposure, improve well performance, and enhance ultimate
hydrocarbon recovery. Figure 1-3 shows how these factors can improve reservoir development
economics and help maximize the return on investment from drilling a well.
• True vertical depth (TVD)—The vertical depth of the wellbore independent of its path
(Fig. 1-4). In the case of a vertical well, measured depth (MD) is the same as TVD.
Figure 1-4. True vertical depth is the vertical depth of the well independent of the well path.
True
Vertical
Depth
(TVD)
Azimuth
Displacement
Figure 1-5. Displacement is the shortest horizontal distance between two points in the
well. It is generally referenced to the surface well location.
• Azimuth—The angle between the displacement line and true north or magnetic north
measured in a horizontal plane, typically measured clockwise from north.
Well position information is generally presented in a 2D format using plan views (looking down
from above) and vertical sections (looking horizontally at a projection of the wellbore in a vertical
plane). Figure 1-6 shows a typical pair of 2D representations of a well trajectory. The planned well
(black line) and the well as drilled (blue line) are shown in side view perpendicular to the planned
well and in plan view from above. In this example the drilled well has departed from the plan in
both TVD and azimuth. Plots such as this are updated regularly while drilling to enable the driller
to compare how close the drilled well is to the well plan.
The plane of the vertical section is generally along the planned well azimuth, but it can be created
at any azimuth. This is demonstrated in Fig. 1-7, where different vertical section lengths are
shown depending on whether the vertical section plane is taken along the planned well azimuth
(green vertical section length) or an alternative azimuth (brown vertical section length). Care must
be taken when looking at vertical section views to ensure that the azimuth of the section is known
because the vertical section length of a well varies depending on the azimuth to which it is
projected. In an extreme case, a horizontal well drilled perpendicular to the section view azimuth
appears as a point, giving no indication of the length of the well.
The true horizontal length (THL) of a well is the length along a projection of a trajectory to the
horizontal plane. THL is independent of well and vertical section azimuths because it follows the
changes in well azimuth.
l
el
W
ll
d
We
ne
an
lled
Pl
Dri
TVD
Dr
ill ed
W
Plan View
ell
Current Survey
•
Planned Well
VERTICAL SECTION
Vertical Section View
Figure 1-6. Well trajectory information is generally presented in vertical and horizontal
projections called the vertical section view and plan view, respectively.
N
ell
• d W
n gth anne
Le Pl
ng
al
lo h
n a ut
ont
tio zim
ec l a
l S wel
oriz
c a
rti ed
Ve lann
True H
p
ng
n alo th
o u
cti zim
l Se ive a
a t
rtic na
Ve alter
an
Plan View
Figure 1-7. The vertical section lengths shown in plan view differ depending on the
azimuth on which the drilled well trajectory is projected. The true horizontal length of the
wellbore is independent of the well and vertical section azimuths.
A well drilled to its final depth is said to have reached total depth (TD).
• Model, compare, and update is the original method, which involves modeling log
responses based on a formation model and well trajectory, comparing the modeled
responses with real-time measured logs, and updating the formation model to match the
real-time measured logs. This method can be applied to any real-time log data.
• Real-time dip determination requires formation data from opposite sides of the wellbore,
preferably images scanned from the entire inner circumference of the wellbore, that are
transmitted while drilling. Formation dip is calculated by the correlation of features from
opposite sides of the borehole. The dip is then extrapolated away from the borehole and
the well steered on the assumption that the formation dip does not change significantly.
• Remote detection of boundaries for real-time well placement currently requires the deep
azimuthal electromagnetic measurements of PeriScope* distance-to-boundary service.
Through an inversion process, the distance and direction to changes in formation
resistivity can be determined. Well placement using this technique requires knowledge of
the resistivity boundaries within a reservoir sequence of layers and which of those
boundaries the measurements and inversion can detect.
Traditional well placement involves specification of a geological target or targets which drilling
engineers then design a well trajectory to intersect and follow. In engineering a planned
trajectory, drilling engineers must consider factors such as avoidance of nearby wells
(anticollision); hole cleaning, removal of cuttings, and formation pressure control (hydraulics); the
ability to manipulate the drillstring without exceeding drillpipe torque and tensile limits (torque and
drag); and the ability to steer the well (bottomhole assembly [BHA] tendencies). Flexibility must
be built into the design of a well to allow for the possibility that the targets could require
modification during drilling. Target locations are generally selected based on a reservoir structural
model derived from seismic data and well-to-well log correlation. Owing to the limited vertical
resolution of surface seismic data and the assumptions made during the well-to-well log
correlations, the actual subsurface structure is often different from that indicated by the model.
The upper panel of Fig. 1-8 shows an example where three targets have been selected and a
well planned to intersect all the targets. However, the formation was not as smooth and
continuous as expected but had subseismic faulting that, if drilled according to the original
geometric plan, would have resulted in the well having little exposure to the reservoir. Through
the monitoring of real-time data and application of well placement principles, the actual well (red
line in the lower panel) achieved greater reservoir exposure and hence production than planned.
The
TheReality:
Reality:
Figure 1-8. Traditional well placement involves drilling geometrically through targets
specified by geologists, but the geology is often different from expectations.
It is important to remember that all data has associated uncertainty. The limited vertical resolution
of surface seismic data and the assumptions made during well-to-well log correlations result in
formation models with uncertainty in the TVD of formation tops and lateral changes in both
formation dip and thicknesses. Even subsurface measurements such as the position of the
wellbore computed from surveys have associated uncertainty. Each survey station has an
ellipsoid of uncertainty associated with the uncertainty in the MD of the survey and the accuracy
of the magnetometers and inclinometers used for the survey measurements themselves. Further
information on surveying uncertainty can be found in Chapter 4, “Measurement-While-Drilling
Fundamentals.” Because the position of a well is computed from these surveys, the accumulation
of uncertainties from each of the survey stations results in an expanding cone of uncertainty
along the path the well is expected to follow with a given probability (Fig. 1-9).
Zone of Interest
Figure 1-9. The cone of uncertainty of the well location results from the accumulation of
uncertainties from each of the survey stations used in calculation of the well location.
The model, compare, and update technique requires that a structural model of the formation of
interest is built and populated with the formation parameters that will be measured in real time.
These formation parameters are typically gamma ray (GR), resistivity, density, and neutron
response. The real-time measurements are made with logging-while-drilling (LWD) tools and the
data transmitted to surface in real time using an MWD system. The MWD surveys are used to
define the position of the well trajectory in 3D space and the corresponding point on the formation
model. With the LWD tool position and the properties of the surrounding layers known, the
theoretical response of the LWD measurement can be computed and compared with the
measured response from the downhole tool. If they match, then the model gives a reasonable
representation of the borehole position relative to the surrounding layers. If they differ, then the
formation model must be adjusted to give a match between the theoretical LWD response
computed from the model and the measured response from the downhole LWD tools.
Formation Property Data (GR, The offset well log data The markers (black) are linked to surfaces (blue) in the
density, neutron, resistivity) (green) is squared (red) to geological model. The squared logs values are propagated
from a representative offset define the layer boundaries across the curtain section between the surfaces.
well is obtained. and formation markers
(black) are identified
Figure 1-10. Construction of the property model involves squaring the offset well logs and
propagating the formation properties into a geological structure model.
The geological structure is generally derived from surface seismic data and well-to-well log
correlation. Faults, formation dip, or changes in layer thicknesses are indicated by changes in the
surfaces (blue lines in Fig. 1-10), which represent the tops of geological sequences. The markers
in the layer column are associated with surfaces in the geological model and the formation
property is propagated into the geological model between the surfaces.
The distribution of the properties between the surfaces depends on the geological deposition of
the layers. Generally three options are available (Fig. 1-11): proportional propagation,
propagation parallel to the upper surface, and propagation parallel to the lower surface. The
selection of the propagation method should be based on the way the subsurface layers were
actually deposited. Proportional propagation is the most common method.
100 degC
90 degC
?
?
100 degC
Inversion of logging tool responses is more complicated than the water temperature example.
Factors such as vertical and radial volumes of investigation, incidence angle between the
wellbore and formation layers, as well as environmental complications such as borehole and
invasion effects must be taken into account to derive the true formation properties.
Note that because the geological model is populated with the formation properties and uses the
planned well trajectory, the forward-modeled log responses reflect what the real-time logs will
look like only if the geology is as mapped and the trajectory is drilled exactly as planned (Fig. 1-
12).
0
A40H - Attenuation
P40H – Phase
1000
Resistivity
(ohm-m)
100
10
2450
GR
(GAPI)
2450
True Vertical Depth (ft)
2450
2450
2450
2450
2000 3000 0 200
True Horizontal Length (ft) GR (GAPI)
Figure 1-12. The forward-modeled log response along the planned trajectory in the
expected formation displays the perfect case only where the geology and well are exactly
as expected.
To be able to recognize deviations from the plan, alternative scenarios should be modeled. For
example, log responses as the well exits the reservoir layer through the roof (top of the layer) and
the floor (bottom of the layer) should be modeled so that these situations can be recognized
during drilling. Both the structural map and offset well log data should be available for reference in
case the well encounters any unexpected features, such as faults or dip changes.
It is important that the operating company’s well placement objective is clearly defined and
understood, that the formation is modeled as accurately as possible, and that likely alternative
scenarios are investigated. During this prejob modeling phase it is the responsibility of the well
placement engineer to ensure that the selected LWD measurements are sufficient to achieve the
well placement objective. If there is a need for additional measurements (for example, quadrant
data or real-time images for formation dip determination), then the value of these measurements
and the well placement limitations without them must be recognized early enough to provide
sufficient time to make the necessary adjustments to the acquisition program.
Figure 1-13 shows an example of a prejob formation model with the LWD logs forward modeled
along the planned trajectory. The three horizontal log tracks at the top of the panel show the
modeled GR (green curve), phase resistivity (blue curve), and ring resistivity (black curve). The
lower panel shows a vertical section along the planned well trajectory on which the formation GR
property model is shown with dark colors representing layers with high GR values and light colors
(GAPI)
GR
0
1000
P40H - Phase
Resistivity
(ohm-m)
100
10
1000
Resistivity
(ohm-m)
100
Ring
10
GR 2450
(GAPI)
True Vertical Depth (ft)
2450
2450
2450
2450
2450
2000 3000 0 200
Figure 1-13. At the end of the prejob modeling phase, a structural model populated with
the formation properties from an offset well has been built and the theoretical log
responses forward modeled for the proposed LWD tools along the planned trajectory.
Once the data is transmitted to surface it must then be distributed to the well placement team.
This is typically achieved by uploading the data to a secure, Web-enabled distribution service.
With the data available to the decision makers, it must be visualized and compared with the
modeled responses to evaluate any deviations from the plan. Software such as RTGS* Real-
Time GeoSteering combines the ability to create and modify a formation model with real-time
data streaming and image and log forward modeling along with the ability to display and compare
all this information in real time to enhance well placement decision making.
The real-time trajectory and log data are streamed into the forward-modeling software usually via
a Web data-delivery service. Initially the formation model remains the same, but because the real
trajectory may differ from the planned trajectory the forward-modeled logs are recomputed along
the real trajectory. In this stage the measured and modeled logs are compared. If they show a
good match, this indicates that the formation model and the location of the real trajectory within in
the formation are representative of what is occurring downhole (Fig. 1-14). Where the measured
and modeled logs do not agree (Fig. 1-15) suggests that the formation model is not
representative of the subsurface and must be updated. The structural model can be changed by
adjusting the dip of surfaces or layer thicknesses or by inserting faults.
Typically in well placement using the model, compare, and update technique, the first formation
model adjustment is a vertical (TVD) shift to tie in a clear formation feature with the depth at
which it is observed in the real well. No further TVD shifts are permitted from this point forward
(unless a fault is inserted in the model) because they would disrupt the original tie-in correlation.
The convention is to work from left to right across the model in the direction of increasing
measured depth. Each new correlation creates either a change in the local dip of the formation or
a change in the layer thickness. In the absence of any information about layer thickness changes,
the formation dip is generally adjusted.
Real-
Real-Time Modeling
200
(GAPI)
GR - measured (GAPI)
GR
GR - modeled (GAPI)
0
1000
P40H - Phase
Resistivity
(ohm-m)
1000
Resistivity
(ohm-m)
GR 2450
(GAPI)
True Vertical Depth (ft)
2450
2450
2450
2450
2450
1000 1500 0 200
Figure 1-14. During real-time well placement operations, measured data is compared with
the modeled logs (red curves). Where they show a good match, as in this example,
Real-
Real-Time Modeling
(GAPI) 200
GR - measured (GAPI)
GR
GR - modeled (GAPI)
0
1000 The actual
P40H - Phase
Resistivity
(ohm-m)
GR 2450
(GAPI)
True Vertical Depth (ft)
2450
2450
2450
2450
2450
1000 1500 0 200
Figure 1-15. When there are differences between the measured and modeled logs, as
shown by the divergence of the log curves from the model (red curves) in the three top
tracks, the formation model must be updated to more accurately reflect the subsurface.
Figure 1-16 demonstrates how an increase in formation dip can be used to adjust the match
between the measured and modeled logs. In this case, the formation dip was increased slightly
so that the real trajectory now lies in the lower part of the lighter colored reservoir section, rather
than in the middle as shown in Fig. 1-15. When the log responses are modeled with the well
traversing the lower layers in the reservoir, the match between the measured and modeled logs is
much improved, indicating that it is a good representation of the well in the subsurface layers.
Note that Fig. 1-16 shows the well toward the bottom of the reservoir interval; it should be steered
up to stay in the middle of the target reservoir section.
(GAPI)
GR - measured (GAPI)
GR
GR - modeled (GAPI)
0
1000
P40H - Phase
Resistivity
(ohm-m)
1000
Resistivity
(ohm-m)
GR 2450
(GAPI)
True Vertical Depth (ft)
2450
2450
2450
Figure 1-16. Adjusting the modeled formation dip by increasing it slightly from that shown
in Fig. 1-15 results in the well trajectory crossing formations that give a better match to the
measured logs.
This iterative procedure of modeling the tool responses, comparing them with the real data, and
updating the formation model so that they match is repeated as the well is drilled to determine the
position of the well trajectory in the layers and make changes to it to ensure that the well stays in
the target layer.
Figure 1-17 shows the final well location on the updated reservoir model. Note the additional 400
ft [120 m] of reservoir exposure in comparison with the geometrically drilled well shown in Fig. 1-
18.
(GAPI)
GR - measured (GAPI)
GR
GR - modeled (GAPI)
0
1000
P40H - Phase
Resistivity
(ohm-m)
1000
Resistivity
(ohm-m)
GR 2450
(GAPI)
True Vertical Depth (ft)
2450
2450
2450
2450
2450
2000 3000 0 200
Figure 1-17. Evaluation of the well location within a reservoir layer enhances both
placement of the well in real time and understanding of the formation structure.
(GAPI)
GR - measured (GAPI)
GR
GR - modeled (GAPI)
0
1000
P40H - Phase
Resistivity
(ohm-m)
1000
Resistivity
(ohm-m)
GR 2450
(GAPI)
True Vertical Depth (ft)
2450
2450
2450
2450
2450
2000 3000 0 200
Figure 1-18. Drilling the well geometrically according to the original plan would have
delivered less reservoir exposure and production.
Making appropriate adjustments to the trajectory as it is being drilled can keep the well in the
target layer, thereby increasing reservoir exposure and hence production. In addition, knowledge
of the structure of the reservoir is improved. This information can be used in refining the 3D
reservoir model for subsequent reserves calculations, production simulation, and future well
planning.
Figure 1-19 shows three different scenarios that could explain a decrease in average density
porosity, each of which requires a different well placement response.
†
The drilling perspective is referenced to the wellbore, so layers approach and cross or traverse
the wellbore, although in reality the situation is the opposite.
Figure 1-20 shows examples of quadrants and sector measurements. Note that the density and
photoelectric images are subdivided into 16 sectors, whereas the density and photoelectric
measurements are delivered as quadrants. This is because these statistical measurements
require counts from four sectors to have the statistical precision required for use as quantitative
measurements. Similarly, the laterolog image is delivered with 56 sectors whereas the resistivity
measurements are in quadrants because of signal-to\-noise considerations.
Up
Left Right
Bottom
Figure 1-20. The azimuthal resolution of each measurement depends on how tightly the
measurement can be focused. Because quantitative measurements generally require
averaging across several sectors to improve measurement signal-to-noise, the image is
delivered in sectors but the measurement can be in quadrants.
The vertical resolution of images is related to how tightly the measurement can be focused.
Figure 1-21 compares the pixel sizes of various LWD and wireline imaging technologies. The
vertical resolution of the LWD images is influenced by the drilling ROP compared with the
sampling rate of the imaging measurement. For example, when drilling at 36 ft/h [11 m/h], the
LWD tool moves 1 ft [0.3 m] in 100 s. If the measurement cycle takes 10 s, then 10 samples per
foot, or 1 sample each 1.2 in [3 cm], are acquired, setting the lower limit to the vertical resolution.
Formation
Ultrasonic Micro
One Borehole Imager
inch GeoVISION
scale Laterolog Imager (FMI)
Resistivity (UBI)
(GVR)
Azimuthal Density
(ADN)
Figure 1-21. Comparison of the relative pixel size of common LWD and wireline imaging
technologies in a 6-in [15-cm] borehole. Each pixel represents the area of the borehole
wall resolved. From the scale bar on the left: LWD density image pixel, 16 sectors; LWD
laterolog image pixel, 56 sectors; wireline UBI* Ultrasonic Borehole Imager pixel; and
wireline FMI* Fullbore Formation MicroImager pixel.
Azimuthal measurements detect changes in formation properties around the borehole. Images
made with different measurement physics from the same borehole can show different features
because the formation properties they measure do not change in the same way. Figure 1-23
shows five LWD images acquired at the same time in the same borehole. The acoustic image on
the far left responds to tool standoff and borehole breakouts; the photoelectric image responds to
changes in formation lithology; the density image responds to changes in formation lithology,
porosity, and fluid content; the GR image responds to the total formation gamma ray count (from
thorium, uranium, and potassium); and the resistivity image responds to the porosity and water
content of the pore space. Although the images show common events that can be correlated, the
differences between them yield additional information about the formation.
100 ft
Figure 1-23. The various LWD images respond to different formation properties and hence
display different formation features.
It is the formation structural information contained in azimuthal data and images that is of interest
for well placement. When the borehole crosses a layer with some contrast in formation property
relative to the zone being drilled, the azimuthal measurements detect the layer as it is traversed
from one side of the borehole to the other. The 2D LWD images are generally displayed as if the
borehole had been split along the top and unfolded so that the center of the image corresponds to
the bottom of the borehole (Fig. 1-24).
When a borehole is drilled down into a layer, the bottom of the borehole sees the layer first, then
the sides see it, and finally the top does. On the image the layer appears first in the middle and
then creates a sinusoidal shape that ends at the edges of the image, corresponding to where the
last of the layer is seen at the top of the borehole. The amplitude of this sinusoid is related to the
incidence angle between the wellbore and the layer. Even without calculation of the incidence
angle, the “happy face” and “sad face” features on an image that result respectively from drilling
up through layers and down through layers can assist in well placement decision making because
they indicate where the wellbore is positioned relative to the layering. For example, if happy face
features are seen on the image while drilling in a reservoir, this indicates drilling up through the
sequence, and a drop in well inclination may be required to become parallel to the layering and
remain in the reservoir.
Because it is responding to the local geometry of the layering with respect to the tool, a happy
face feature on a borehole image always indicates that the well is penetrating up into a layer.
Similarly a sad face feature always indicates that the well is drilling down into the top of a layer,
irrespective of the trajectory, layer dip or azimuth.
As the incidence angle between the borehole and layer decreases the amplitude of the sinusoid
on the image increases. This means that in a high angle well through horizontal layering a thin
bed appears stretched over significantly longer measured depth than in a near vertical well.
For example, in an 8.5-in [21.6 cm] well drilled at an incidence angle of 1° through a 6-in– [15-
cm-] thick bed, the bed will appear on the image stretched over more than 69.2 ft [21.1 m] of
measured depth. Refer to Fig. 5-4 for a sketch of this geometry. In a well drilled perpendicular to
the layer the 6-in– [15-cm-] thick bed would not be fully resolved by a measurement with an axial
resolution of 1-ft [30 cm]. However in a well close to parallel with the bed even a 1-ft [30 cm] axial
resolution measurement will be able to identify the feature as it is stretched over 69.2 ft [21.1 m].
This geometrical stretching of formation features in high angle wells greatly enhances the visibility
of thin layers. Consequently even layers that are too thin to be fully identified in near vertical wells
can be used for well placement purposes in high angle wells.
Image normalization is a method by which features can be visually enhanced (Fig. 1-25). A static
image has a fixed, user-defined scale over which the color spectrum is applied. For example, the
3
colors of a density image may be scaled over 16 colors, with dark colors starting at 2.2 g/cm to
3
light colors at 2.7 g/cm . A dynamic image uses a depth window within which the minimum and
maximum values are used as the endpoints for the color scaling. This method allows subtle
contrasts between features to be enhanced.
Statically
normalized
image
Dynamically
normalized
image
Figure 1-25. In these statically (upper) and dynamically (lower) normalized images over the
same interval, the static image has a fixed density range for each color. The dynamic
image uses the full range of available colors over a user-specified depth window.
It is generally good practice to use both static and dynamic images because the static image
highlights large-scale features whereas the dynamic image enhances features within the
normalization depth window, which can reduce the visibility of the large-scale features.
During rotation of the LWD tool about its axis, the sensors sweep out a circle in the plane
perpendicular to the axis of the tool. The orthogonal magnetometers, mounted perpendicular to
the axis of the tool, measure the projection of the Earth’s magnetic vector in this plane. The
projection of the Earth’s gravitational vector in this plane defines the direction of the bottom of the
hole. Data acquired relative to the projection of the magnetic vector must be rotated through the
angle shown in red on Fig. 1-27 to orient it relative to the projection of the Earth’s magnetic
vector.
_X
Angle
Figure 1-27. Azimuthal data is initially oriented relative to the projection of the Earth’s
magnetic vector on the plane perpendicular to the axis of the tool. The data must be
rotated to orient it to the top-of-hole reference, which is the standard orientation in high-
angle and horizontal wells.
Because the orientation of azimuthal data requires a projection of the magnetic field on the plane
perpendicular to the axis of the tool, care must be taken when drilling close to parallel to the
Earth’s magnetic vector. If the well is drilled along the magnetic vector there is no projection of
the magnetic field detected by the orthogonal magnetometers in the tool (Fig. 1-28). In this case
azimuthal data cannot be oriented to either a north reference or the top of hole.
Figure 1-28. A zone of exclusion exists around the Earth’s magnetic vector where the tool
and well azimuth cannot be determined.
Within this cone around the Earth’s magnetic field the amplitude of the magnetic field sinusoids
detected by the orthogonal magnetometers remains too small to reliably orient azimuthal data.
During well planning a zone of exclusion is defined around the Earth’s magnetic vector in which
azimuthal data cannot be acquired. Wells are generally planned with the minimum possible
interval in the zone of exclusion.
• Adjacent side—The amplitude of the sinusoid on the image is measured along the
borehole. The measured depth amplitude of the sinusoid must be converted to the same
units as the borehole diameter.
• Opposite side—Images do not scan the surface of the borehole. They represent the
formation property at the depth of investigation of the measurement from which they are
derived. Therefore, the depth of investigation of the imaging measurement must be
added on each side of the borehole diameter. For density measurement, the depth of
investigation is approximately 1 in [2.5 cm], so for an 8.5-in [21.6-cm] borehole the
diameter at which the image is acquired is 10.5 in [26.7 cm]. For laterolog resistivity
image the electrical penetration depth is approximately 1.5 in [3.8 cm], so 3 in [7.6 cm]
should be added to the diameter of the borehole to calculate formation dip from an LWD
resistivity image.
Figure 1-29 shows a wellbore crossing a layer with the corresponding density image and
incidence angle calculation.
Density
image
bore
hole
8.5+ 612
2 =10. ”
5”
layer
Sinu
BH + 2 x DOI soid
a mpli
= 8.5” + 2 x1” tude
=10.5” in M
D=
612
”
α
α = tan–1(10.5/612)
= 1° between the borehole and layer
Figure 1-29. The incidence angle between the borehole and formation layering can be
deduced by solving the trigonometry of a triangle. BH = borehole, DOI = depth of
investigation.
The plane in Fig. 1-30 has a strike northeast–southwest (45°) and a dip of 30° degrees at an
azimuth of 135° degrees (southeast).
Figure 1-30. Strike is the azimuth of the intersection of a plane, such as a dipping bed, with
a horizontal surface. Dip is the magnitude of the inclination of a plane from horizontal.
True, or maximum, dip is measured perpendicular to strike.
Well azimuth
Figure 1-31. The projection of formation dip on a vertical plane is at a maximum when the
vertical plane is oriented perpendicular to the strike, giving true dip. Projection in a
vertical plane along any other azimuth has a lower dip value, the apparent dip.
Apparent dip refers to the projection of the true dip in an azimuth other than perpendicular to
strike. In a well placement context, the more significant angle is the incidence angle between the
borehole and the layer (Fig. 1-32).
To remain parallel to a layer while drilling, the incidence angle is all that is required to make
trajectory change decisions, provided that the well continues to be drilled in the same azimuth. If
the well changes azimuth (that is, turns left or right) then the incidence angle will change, and this
must be taken into account to be able to stay in the target layer.
Boreh
The borehole is in the plane of the page.
ole
Figure 1-32. The incidence angle is measured between the wellbore and the formation
layer. Apparent dip is the angle between the wellbore and the horizontal.
Calculating the true dip of the formation from the incidence angle requires that the borehole
trajectory be taken into account. This is because the incidence angle is measured by the logging
instruments from the perspective of the borehole. For example, the incidence angle between a
vertical borehole and formation layers dipping at 45° is the same as the incidence angle between
a 45° inclined borehole and horizontal layers.
1.5.2.9 3D visualization
Real-time 3-D visualization and dip-picking software is available to assist in visualizing the
intersection of formation layers and inclined boreholes. By wrapping an oriented image around a
3D representation of the corresponding well trajectory, 3D visualization software helps the user
understand the subsurface orientation of the wellbore and layers that it intersects.
Figure 1-33 shows a WellEYE* 3D borehole data viewer display. The right panel is a conventional
2D display of laterolog resistivity log and image data. The green sinusoid on the image tracks is
where a geological feature has been picked. The corresponding formation dip information is
indicated by the green tadpole in the dip track between the two image tracks. The left panel
shows a 3D perspective of the well trajectory with the laterolog image data “wrapped” around the
The ability to transmit real-time image data from downhole and stream it into interactive
visualization and dip-picking software, either at the wellsite or through an Internet connection,
enhances the value of the information because it enables almost instantaneous quantitative
interpretation to improve decision making.
Figure 1-33. Real-time 3D visualization and dip-picking software enhances the user’s
understanding of the subsurface geometry.
The example in Fig. 1-34 is a typical LWD log from a horizontal borehole. The top track shows
quadrant density responses and an azimuthal average thermal neutron response. The middle
track shows three phase resistivities of differing depths of investigation. The depth track includes
a pink line labeled ARPM (adnVISION revolutions per minute) that indicates where the density
There is a 4-m [13-ft] MD difference between the density drop seen by the bottom and up density
measurements. The bottom density measurement decreases first, which indicates that the well is
being drilled down into a layer of lower density. The TVD curve shows that the well is horizontal,
so the layer must dip upward.
Well TVD
ROP
GR
Figure 1-34. This typical LWD log from a horizontal borehole displays quadrant density
information in the top track. The bottom density (red curve) and up density (green curve)
show a clearly correlated drop in density.
For this 6-in- [15.24-cm] diameter well, the calculation of incidence angle (α) is as follows:
MD difference = 4 m = 157.5 in
Borehole diameter = 6 in
Density measurement depth of investigation = 1 in
α
–1
= tan 0.0508
= 2.9° incidence angle between the borehole and la yer.
The incidence angle calculation, derived either from images or quadrant curves, is used to
position the wellbore relative to the layer. For the example in Fig. 1-34, if the intention is to
Top of layer
Me
as
u
Th Dept red
ick h
ne
ss
TVDT – True
Vertical Depth
Thickness dip
TBT - Tru
eB
Thickne ed
ss
Bottom of layer
Figure 1-35. The true bed thickness cannot be determined from conventional logs unless
the formation dip is known.
In the absence of formation dip information, the conventional model, compare, and update well
placement method generally assumes that a formation layer thickness is the same as observed in
an offset well (Fig. 1-36).
Calculated Assumed
Dip of the TBT - True Bed
layer Thickness
Incidence angle calculated
Unknown
Incidence # 1to the
angle Assume
Unknown
that the
# 2true
layer can then be thickness of the
calculated layer is constant
Figure 1-36. Formation dip can be calculated by assuming that the layer thickness in the
drilled well is the same as in an offset well.
The assumption of constant layer thickness can result in an invalid formation model and poor well
placement. Nonazimuthal data cannot distinguish between change in the dip of the layer, change
in the thickness of the layer, or change in both. A combination of layer thickness and dip can
explain any apparent MD layer thickness (Fig. 1-37).
Top of layer
Dip of the TBT - True Bed
layer Thickness
Unknown # 1 Unknown # 2
high incidence angle thick layer
Figure 1-37. Various combinations of layer thickness and dip can explain the same
apparent thickness in MD.
By enabling determination of the formation dip at the point where it crosses the borehole,
azimuthal data makes it possible to calculate the TBT (Fig. 1-38).
Drilling down
structure
Calculated Calculated
Dip of the TBT - True Bed
layer Thickness
Known incidence angle
Unknownfrom
Evaluated #1 Unknown
Can now #2
be calculated
the LWD Image without assumptions
Thickness computed
Figure 1-38. Information on the layer dip that is available from azimuthal measurements
enables building a more accurate formation model by removing the need to make
assumptions.
Real-time dip determination from azimuthal data is a powerful, complementary method for use
with the basic model, compare, and update technique for well placement. Azimuthal data
provides information about the direction from which a feature approaches and crosses the
borehole and the dip of the feature, which allows the TBT to be calculated.
l
d
Figure 1-39. An azimuthal measurement is usually located a distance (l) behind the bit.
d = lsin α (2)
where
d = distance to the trajectory perpendicular to the layer (assumes that the layer and trajectory do
not change in angle)
l = length between the azimuthal measure point and the bit
α = incidence angle between the trajectory and layer.
Tilted antennae on the PeriScope tool create directional sensitivity to conductivity changes in the
formation. Directional measurements of electromagnetic phase shift and attenuation are analyzed
downhole to determine the direction to the nearest conductivity contrast and then transmitted
from the downhole tool to the surface, where inversion processing extracts the distance-to-
boundary (DTB) information.
hd
Rd
Figure 1-40. PeriScope inversion processing solves for the resistivity and distance to layer
boundaries above and below the wellbore based on a three-layer model (left). The
resulting information is displayed in real time as a color-coded resistivity cross section of
the formation (right). Ru = resistivity of the upper layer, Rh = horizontal resistivity of the
local layer, Rv = vertical resistivity of the local layer, Rd = resistivity of the lower layer, hu =
distance to the upper layer, hd = distance to the lower layer.
In addition to DTB information, the PeriScope service provides an azimuth to the boundaries,
based on the assumption that the layers above and below are parallel. This information is
presented in an azimuthal view (Fig. 1-41).
The distance and azimuth information enable steering of a well in both TVD and azimuth relative
to a resistivity boundary without having to come in contact with it. For example, in the situation
shown in Fig. 1-41, the well could be either turned up to avoid the lower boundary or turned to the
left, or a combination of the two could be used depending on the most appropriate position for the
wellbore in the target layer.
Distance to Boundary
15 ft
Propagation Resistivity
4 ft
Images
Figure 1-42. PeriScope DTB service delivers both directionality and depth of investigation
at a subseismic level.
Seismic data can visualize large structural features, but owing to the limited frequency content of
surface seismic data, many features are below seismic resolution. PeriScope service, with a
radius of investigation of approximately 15 ft [4.5 m], can be used to evaluate events such as
subseismic faults (Fig. 1-42). As thinner reservoirs are drilled, the evaluation of subseismic
faulting becomes increasingly important because the borehole can exit the reservoir on
encountering a fault. In addition, the evaluation of reservoir compartments is improved through
the ability to track the top and bottom of the layer, thereby improving estimation of the volume of
hydrocarbons in place. Further details on are in Section 5.14, “Remote boundary detection.”
There are a number of directional drilling technologies, such as jetting, whipstocks, and rotary
steerable assemblies. The two most commonly used for well placement are the steerable motor
and rotary steerable system (RSS).
Steerable motors consist of a positive displacement motor (PDM) with a surface-adjustable bent
housing, which enables orienting the bit in the desired drilling direction. PDMs convert hydraulic
power from the mud circulation into mechanical power in the form of bit rotation (Fig. 1-43). . This
is achieved with a progressing cavity design in which the movement of the mud pushes on a rotor
with one less lobe than the stator in which it is housed. The rotor turns inside the stator and thus
turns the bit coupled to the rotor. With a PDM the bit rotates when there is mud circulation, even if
the drillstring is stationary.
Because the angle on the bent housing is small (typically less than 3°) the
steerable motor can also be rotated. This negates the effect of the bend and
gives a relatively straight borehole, which is slightly over-gauge (larger than bit
size). By alternating sliding and rotating intervals the directional driller can control
the rate at which the borehole angle is changed. The rate of change in borehole
angle is normally given in degrees per 100 ft [30 m], and is also called dogleg
severity (DLS).
The difficulty with slide-rotate sequences is that orienting prior to each slide
section is time consuming and hence reduces the overall ROP achieved for the
well. In addition, because the mud is not agitated during sliding, hole cleaning
efficiency is reduced. Finally, the overall length of the well can be limited because
static friction during sliding, which is greater than dynamic friction while rotating,
can prevent effective weight transfer to the bit. To overcome these limitations, the
RSS was developed.
Figure 1-44. The surface-adjustable bent housing is oriented in the direction of the desired
well trajectory.
Rotary steerable systems deliver continuous steering while rotating. This capability beneficially
results in steadier deviation control, a smoother hole, better hole cleaning, extended hole reach,
and overall improvement in the ROP compared with steerable motor performance. It is also
advantageous for well placement because the continuous rotation of the BHA ensures that when
formation images are acquired, they are available over the entire length of the well.
A push-the-bit system uses pads on a bias unit to push against the borehole wall, which pushes
the bit in the opposite direction (Fig. 1-45).
Flex Stabilizer Control Unit Bias Unit
Activated Pad
Figure 1-45. A push-the-bit system uses three pads to push against the borehole wall and
so deflect the well in the opposite direction.
Pad pushing
Steering
direction
Figure 1-46. The spindle (blue) is held geostationary by the control unit, thereby diverting
a small proportion of the mud flow behind each of the pads in sequence as their
respective entry ports rotate in front of the hole in the control valve.
To meet the high power requirements of the control motor, the system generates its own power
by using a high-power turbine and alternator assembly above the control unit.
The dogleg delivered by the push-the-bit system depends both on the interaction of the pads with
the formation (for example, DLS is lower in unconsolidated formations) and the proportion of time
spent steering. The directional driller communicates with tool through a sequence of mud flow
Compared with the slide-rotate sequences of a motor with a bent housing, the continuous
steering of an RSS delivers smoother trajectories, which improves the drilling operation itself, as
well as subsequent casing and completion runs and later well intervention operations (Fig. 1-48).
0
Inclination from continuous survey 0.29 deg./100 ft 1.38 deg./100 ft 4.02 deg./100 ft
Inclination from static survey DLS DLS DLS
Toolface
Figure 1-48. The trajectory oscillations resulting from steerable motor slide-rotate
sequences (left) are eliminated by the continuous steering of an RSS (right).
A point-the-bit system delivers the benefits of a push-the-bit system with reduced sensitivity to the
formation, resulting in more consistent steering and generally higher dogleg capability.
The system is centered on a universal joint that transmits torque and weight on bit, but allows the
axis of the bit to be offset with the axis of the tool. The axis of the bit is kept offset by a mandrel
that is maintained in a geostationary orientation by a counter-rotating electrical motor (Figs. 1-49
and 1-50). To meet its high power requirements, the system generates its own power with a high-
power turbine and alternator assembly. The system also contains high-power electronics to
control the motor and sensors that monitor the rotation of the collar and motor. The sensors
provide input and feedback for control of the system.
Figure 1-49. The PowerDrive Xceed* point-the-bit system uses an electric motor to counter
rotate a mandrel against the rotation of the collar. This keeps the mandrel, and thus the
bit, oriented in the same direction while still rotating with the collar.
As with the push-the-bit system, the directional driller controls the dogleg by downlinking to the
tool to change the proportion of steering versus neutral time.
Because there is no rotation provided downhole by either type of RSS, the entire drillstring must
be continuously rotated from surface. If additional downhole rotation is desired or surface rotation
must be kept to a minimum (such as when casing wear is a concern), a mud motor (without the
bent housing) can be used above the RSS to provide downhole rotation of the RSS assembly.
Knowledge of the well location in a formation and in 3D space is critical for all facets of well
construction and formation evaluation. The inclination of a wellbore from vertical is determined by
using a set of triaxial accelerometers to measure the components of the Earth’s gravitational field.
In conjunction with the inclination data, a set of triaxial magnetometers is used to measure the
components of the Earth’s magnetic field to determine the azimuth of the borehole with respect to
magnetic north. Knowledge of the local horizontal angle between true and magnetic north
(magnetic declination) enables conversion of the azimuth to a true north reference. In many
cases a grid convergence correction is applied to account for local distortion resulting from the
projection of the curved surface of the Earth onto a flat, 2D map. Azimuths corrected for this
effect are said to be referenced to grid north.
During directional drilling, the orientation of the drilling system defines the direction in which the
well deviates. Toolface is the angle between a reference, either gravity in a deviated well or north
in a vertical well. Because the toolface is the direction in which the BHA tends to deviate the hole,
it is used by the directional driller to ensure that the drilling assembly is oriented to give the
desired direction to the well.
Batteries could be used to deliver power to the measurement electronics and telemetry system,
but the duration of drilling runs would be limited to the life of the batteries. For this reason many
MWD systems incorporate a downhole mud turbine and alternator as an electrical power
generation system (Fig. 1-51). Whenever mud is being pumped through the drilling system,
hydraulic power from the mud flow is converted into electrical power as the turbine rotates and
drives the attached alternator. The generated electrical power is then available to the MWD
subsystems and, where an intertool electrical connection is available that provides power and
data connectivity along the BHA, power from the MWD turbo-alternator system can also be used
by other tools in the BHA.
Figure 1-51. Hydraulic power from the mud flow is converted into electrical power through
use of a turbine and alternator system. The stator (blue) deflects the mud flow (from left to
right) on to the rotor (red) causing it to turn. The alternator (inside the yellow housing to
the right) converts the rotation to electrical power.
There are now several methods for transmitting data from downhole tools to the surface including
electromagnetic propagation and wired drillpipe; however, the vast majority of real-time data
transmission from downhole to surface is still performed by mud-pulse telemetry.
Mud-pulse telemetry involves encoding data in pressure pulses that propagate up through the
mud inside the drillpipe. These pressure pulse sequences are detected at surface and decoded to
recreate the numerical value of the data from the downhole tools. There are three main ways of
creating a mud pressure pulse (Fig. 1-52). Positive pulse systems impede mud flow with a poppet
valve, resulting in a temporary increase in pressure. Negative pulse systems use a bypass valve
to bleed pressure off to the annulus, resulting in a temporary drop in pressure. A continuous
carrier wave, or siren, system uses a rotating valve assembly that alternates between opened
and closed positions, resulting in an oscillating pressure wave. Data can be encoded using the
siren system by frequency, phase, or amplitude modulation.
Pressure
Time
Pressure
Time
Pressure
Time
Figure 1-52. The three major mud-pulse telemetry systems are the positive pulse, negative
pulse, and siren system, which generates a continuous carrier wave.
The tools used to acquire formation evaluation measurements while drilling are generally referred
to as logging-while-drilling tools to distinguish them from the drilling-oriented MWD tools. In
general, LWD tools send selected formation evaluation data via an internal tool bus to the MWD
tool for transmission to surface. All data transmitted in this way, along with the corresponding
surface data acquired during drilling, is referred to as real-time (RT) data. Downhole tools also
have memory in which all the measured data (as distinct from the limited selection of data sent in
real time) is stored for retrieval when the tool returns to surface. Data extracted from the tool
memory is referred to as recorded mode (RM) data.
When formation evaluation measurements were first migrated from wireline tools to drill collars,
naturally the key triple-combo measurements of resistivity, density, and neutron response were
the first to be made available in addition to the GR measurement, which was also available from
the MWD tools. Owing to the nature of the drilling environment, in addition to the limitations
imposed by having to fit detectors and electronics in a drill collar, some changes were made to
the configuration of sensors. Because of these necessary changes, LWD tools and wireline tools
measuring the same formation parameter generally have slightly differing raw responses. After
appropriate environmental corrections are applied, any remaining discrepancies are generally the
result of differences in the invasion profile and borehole condition between the early LWD log and
the subsequent wireline log.
The evolution of LWD technology has seen significant improvements in the triple-combo services
and the addition of numerous additional measurements, including imaging of multiple formation
properties, magnetic resonance, sonic and seismic acquisition, elemental capture spectroscopy,
and sigma measurements.
In addition to increasing measurement sophistication, LWD data has broadened in use from
petrophysical evaluation of the formation to real-time measurements for evaluation of the wellbore
location within the layers (well placement) and wellbore stability (geomechanics).
Because wireline measurements are acquired after the hole is drilled, the cost of running the logs
includes rig-time costs for the duration of the logging as well as the logging cost itself. The
benefits of having formation data while drilling, in addition to the significant reduction in rig-time
costs associated with running LWD rather than wireline logs, have resulted in LWD being used to
acquire a significant proportion of openhole formation evaluation data.
The decision to run wireline logs or LWD logs is based on three major factors: rig cost, drilling
risk, and well placement risk (Fig. 1-53). If any one of these is a high priority, then most likely
LWD data will be acquired. If all are low in priority, then wireline data acquisition is most likely the
preferred option.
Figure 1-53. The openhole formation evaluation market is served by LWD and wireline-
conveyed measurements.
In wells with high well placement risk, LWD measurements are required to deliver real-time
information about the formation surrounding the borehole to identify the location of the well in the
geological sequence. On the basis of this information the placement of the well can be optimized
relative to the formation structure.
To determine the location of a well within a geological sequence, measurements that enable the
identification of markers must be available in real time. Markers used for steering are generally
changes in a formation property, such as a high-GR layer just above the target reservoir or a
change in formation resistivity that indicates where reservoir hydrocarbon saturation deteriorates.
To detect these markers LWD tools that measure the formation properties must be available.
Over the years the LWD measurement portfolio has expanded significantly, with the following
formation measurements are now available while drilling:
• gamma ray
• resistivity
− laterolog
− propagation
• bulk density
• photoelectric factor
• neutron response
− thermal
− epithermal
• magnetic resonance
− longitudinal relaxation time (T1)
− transverse relaxation time (T2)
− diffusivity of fluids in the pore space (D)
• acoustic
− compressional
− shear
− Stoneley
• elemental spectroscopy
• thermal capture cross section (sigma)
• formation fluid analysis
− pressure
− optical fluid identification
− samples
• seismic
Ongoing research and engineering efforts continue to enhance both the portfolio of
measurements and the capabilities of existing measurements.
With such a wide selection of measurements available, a great deal of information about the
formation surrounding the wellbore can be extracted in real time to assist in determining the
location of the well in the geological sequence. In conjunction with the various well placement
techniques, these measurements support enhanced wellbore positioning in addition to enhanced
formation evaluation.
Data acquired by the downhole LWD tools is compressed, encoded, and transmitted to the
surface, most commonly through a mud-pulse telemetry system. Owing to the limited bandwidth
of current mud-pulse systems (typically 0.5 to 12 bits per second [bps]), the amount of data that
can be transmitted to surface in real time is limited. Improvements continue to be made in the
telemetry rate (the number of bits that can be transmitted per second) and data compression (the
amount of data transmitted per bit). Selection of what data is to be sent is still required because
bandwidth capable of transmitting all the data all the time, as is the case with wireline tools, is
unlikely to be widely available in the near future.
Real-time data is grouped into data points (d-points), each of which represents a particular
measurement (for example, formation bulk density) or is part of a larger collection of data, such
as part of an image that is spread across several d-points.
The d-points are grouped into frames, which define the data to be transmitted when the BHA is in
a particular mode of operation (Fig. 1-54). For example, a frame designed for use when the BHA
is sliding in a deviated well would contain GTF and nonazimuthal formation measurements
because there is no point in wasting bandwidth by sending azimuthal data, which cannot be
acquired when the BHA is not rotating. A frame designed for use when the BHA is rotating would,
in contrast, most likely contain d-points for azimuthal and perhaps image data, but would not
contain a toolface d-point because the toolface is not of interest when the BHA is rotating.
Frame
Synchronization word
and Frame Identifier
d-points
Figure 1-54. The d-points containing measurement data are grouped into frames.
The selection of which frame to transmit is made by the downhole tool based on the well
inclination (MTF or GTF frame), whether the tool is rotating (rotary frame), and whether the mud
flow has just started again after a period of no flow (utility frame). After a few training bits are sent
to allow the surface system to synchronize, the frame identifier is sent. Because both the surface
and downhole systems have been programmed with the same frames, the subsequent stream of
bits is divided into the corresponding d-points and decoded by the surface system.
d-points
Figure 1-55. The bit stream encoded in the mud-pulse pressure wave is divided back into
the d-points by the surface system, which is programmed with the same frame information
as the downhole tool.
Decoding involves the conversion of the binary bit stream to decimal followed by application of
the reverse transform that was applied to the data downhole (Fig. 1-55). For example, a
3 3
downhole density measurement of RHOB = 2.4 g/cm could have 0.9 g/cm subtracted and the
remainder divided by 0.01 to give a decimal number of 150. The eight-bit binary equivalent,
10010110, is transmitted via the mud-pulse system to the surface, where it is converted back to
3
the decimal number 150 and the reverse transform, RHOB_RT = 0.01X + 0.9 g/cm , is applied.
3
The real-time RHOB_RT = 2.4 g/cm measurement (the “_RT” suffix designates it as real-time
1.6.2.1.2 Real-time transmission from the surface system to the decision maker
If the real-time data user and decision maker are at the wellsite, then further transmission of the
data from the acquisition system may not be immediately necessary. However, with the
increasing use of remote operations support, it is likely that the data needs to be securely
distributed to approved users. Data encryption and satellite transmission enable securely
transferring the data from any rig to a satellite receiving station, from which it can be distributed
over the Internet while still encrypted. Dedicated user accounts with password protection ensure
that the data is available only to approved users.
In a few seconds, data acquired under the extreme conditions of downhole drilling can be made
available to approved users anywhere in the world for subsequent interpretation and decision
making.
Once transmitted from the downhole tools to the surface and from the surface acquisition
computer to the decision maker, the data must be presented in a manner that helps the decision
maker extract the relevant information encoded in the data stream.
In the case of image data this is generally best achieved by 2D and 3D visualization of the data,
color coded to represent the measured formation parameter. The addition of interactivity and dip-
picking to the visualization environment allows the decision maker to extract quantitative
information about the formation dip from the data stream. This capability facilitates well placement
using the real-time dip determination technique.
The model, compare, and update method for well placement requires more sophisticated
software support because the incoming real-time data must be displayed compared with modeled
tool responses. The software must be able to create and modify a formation structural model
populated with multiple formation properties and a planned well trajectory. In addition, the
software must be able to simulate (forward model) the response of the LWD tools and stream in
the real-time trajectory and logging data so that they can be compared with the simulated log
response. As discussed in Section 1.5.1, “Model, compare, and update,” discrepancies between
the simulated and real data are an indication that the formation model does not accurately
represent the subsurface formation and hence the model needs to be updated so that the
simulated and actual data match. Once they match, the position of the wellbore in the formation
can be assessed and appropriate well placement decisions taken and communicated to the
directional driller.
Presentation of this data is generally of the form shown in Fig. 1-56. A curtain section of the
formation along the planned well trajectory, color coded to show one of the formation properties
of interest, is plotted in the lower panel with TVD as the vertical axis and THL as the horizontal
axis. Formation structural information is captured in the scaled geometry of the curtain section
layers and faults. The formation model can be color coded for any of the formation properties
entered during the predrilling model construction.
Usually both the planned and real-time well trajectories are displayed so that any departure of the
well from the planned well trajectory can be identified. Note that the curtain section is constructed
during the predrilling preparation phase when only the planned well is available. Hence, the
curtain section is typically a 2D vertical slice through a 3D formation model along the planned well
trajectory. If the actual well trajectory departs significantly from the azimuth of the planned well
trajectory, the curtain section may need to be reconstructed along the new well trajectory. More
Horizontal log tracks in the upper panel display both the real-time data and forward-modeled log
responses so that discrepancies between them can be identified. Image data can also be
displayed in the horizontal log tracks. If a discrepancy between the forward-modeled and real-
time logs is identified, then interactive adjustments are made to the formation model. The
simulated log responses are then recomputed for the edited formation model and compared
again with the real-time data. This model, compare, -and update cycle is repeated until a match is
achieved.
100
(GAPI)
GR
GR - measured (GAPI)
GR - modeled (GAPI)
0
20
Phase Resistivity
(ohm -m)
0.2
RHOB (g/cc)
2.7
Figure 1-56. Forward-modeled and real-time results are compared in the horizontal log
tracks. The position of the wellbore relative to layering in the property model is displayed
in the lower panel.
Because operating company personnel use information that is often accessible only within their
work environment, well placement is best performed in the operating company office. This not
only ensures that the additional information is readily available, but also facilitates communication
and helps the well placement team operate with mutual understanding of the objectives and
operational constraints.
Well placement services provided by service companies have been misconstrued by some
operating company personnel as an attempt to encroach on their responsibilities or even as a
threat to their job security. This is certainly not the intention, because operating company
geoscientists are an integral part of the well placement team whose understanding of the asset
and corporate objectives is critical to ensure that well placement delivers value to the operating
company.
Consider the example discussed previously, in which the deep, directional electromagnetic
measurements indicate a layer of higher resistivity above the wellbore. The well placement
engineer has the technical knowledge to perform quality control on the measurements and
inversion results in consideration of the capabilities and limitations of the service. By presenting
the two possible interpretations of the observed responses (higher resistivity above the well
caused by either lower formation porosity or higher hydrocarbon saturation), the well placement
engineer distills the technical information and suggests possible interpretations. Without requiring
a detailed understanding of the service provider’s technology, the required information is
available for the operating company personnel to apply their skills. Within the operating company
this ensures that the limited expertise on the asset is focused where it adds greatest value. Such
collaboration is the key to successful well placement.
An essential part of prejob preparation is the definition of roles and responsibilities for all involved.
This includes operating company personnel, well placement engineers, directional drillers, and
other wellsite personnel. Names, contact details (e-mail, office and mobile telephone numbers,
fax numbers), responsibilities, and shift times should be documented and distributed to all
involved prior to the start of the job.
The final decision maker must be clearly identified to all team members and contingencies
planned should additional decision makers need to become involved, such as if the well needs to
be abandoned or extended because of poor reservoir quality.
A record must be kept of all decisions and the time at which they were made. Secure Web
communication chat facilities are an excellent communication tool because the author and time of
each entry are logged. This is a valuable record to have if a review of operations is required.
Ultimately, team coordination is about ensuring that the information required to achieve the well
objectives is available to the decision makers and the personnel responsible for executing the job.
In placing wells within a geological target it is important that the well placement engineer have a
basic understanding of the geological processes that create reservoirs. This knowledge not only
enables the well placement engineer to anticipate the most likely shape of the reservoir (for
example, whether layers are straight or undulating) but also be able to discuss possible
geological scenarios with the operating company geologist.
2.1.1 Igneous
Igneous rocks crystallize from molten Earth material, called magma, or lava when it is at the
Earth’s surface. Igneous rocks that crystallize slowly, typically below the surface of the Earth, are
intrusive igneous rocks and have large, interlocking mineral crystals (coarse grained enough to
see with the naked eye). Extrusive igneous rocks crystallize quickly at the Earth's surface and
have small crystals (usually too small to see without magnification) or form a noncrystalline glass.
Igneous rocks typically consist of the minerals quartz, mica, feldspar, amphibole, pyroxene, and
olivine.
2.1.2 Metamorphic
Metamorphic rocks form from the alteration of preexisting rocks by changes in ambient
temperature, pressure, volatile content, or all of these. Such changes can occur through the
activity of fluids in the Earth and movement of igneous bodies or regional tectonic activity. The
texture of metamorphic rocks can vary from almost homogeneous, or nonfoliated, to foliated
rocks with a strong planar fabric, or foliation, produced by the alignment of minerals during
recrystallization or by reorientation. Common metamorphic rocks are slate (metamorphosed
shale) and marble (metamorphosed limestone). Graphite, chlorite, talc, mica, garnet and
staurolite are common metamorphic minerals.
2.1.3 Sedimentary
Sedimentary rocks are formed at the Earth's surface through the deposition of sediments derived
from weathered rocks, biogenic activity, or precipitation from solution.
• Clastic sedimentary rocks such as conglomerates, sandstones, siltstones, and shales
form as older rocks weather and erode and their particles accumulate and lithify, or
harden, as they are compacted and cemented.
• Biogenic sedimentary rocks form as a result of shell building or other skeletal growth of
organisms, such as coral reefs that accumulate and lithify to become carbonate, primarily
limestone and dolomite.
• Precipitates form when solids concentrate as water is lost to evaporation. The evaporite
minerals halite (salt) and gypsum can form vast thicknesses of rock as seawater
evaporates.
Sedimentary rocks can include a wide variety of minerals, but quartz, feldspar, calcite, dolomite,
and evaporite group and clay group minerals are most common because of their greater stability
at the Earth's surface than that of many minerals in igneous and metamorphic rocks. Sedimentary
The dynamic nature of the Earth can change rocks from one type to another. Deep burial and
melting create igneous rocks from preexisting rocks and sediments. Deformation and
metamorphism produce metamorphic rocks from preexisting rocks and sediments. Uplift, erosion,
and deposition can form sedimentary rocks from preexisting rocks and sediments.
Figure 2-1. The rock cycle changes rocks from one type to another as a result of the
dynamic nature of the Earth.
Alluvial deposition pertains to the subaerial (as opposed to submarine) environment, action, and
products of a stream or river on its floodplain, usually consisting of detrital clastic sediments. It is
distinct from subaqueous deposition, such as in lakes or oceans, and lower energy fluvial
deposition. Sediments deposited in an alluvial environment can be subject to high depositional
energy, such as fast-moving flood waters, and may be poorly sorted or chaotic.
Lacustrine deposition pertains to deposition in lakes or an area with lakes. Because deposition of
sediment in lakes can occur slowly and in relatively calm conditions, organic-rich source rocks
can form in lacustrine environments.
Eolian deposition pertains to the deposition of sediments by wind, such as sand dunes in a
desert. Because fine-grained sediments such as clays are removed easily from wind-blown
deposits, eolian sandstones are typically clean and well sorted.
Fluvial deposition pertains to deposition by a river or running water. Fluvial deposits tend to be
well sorted, especially in comparison with alluvial deposits, because of the relatively steady
transport provided by rivers.
Deltaic deposition pertains to an area of deposition or the deposit formed by a flowing sediment-
laden current as it enters an open or standing body of water, such as a river spilling into a gulf. As
a river enters a body of water, its velocity drops and its ability to carry sediment diminishes,
leading to deposition. Sediments characteristically coarsen upward in a delta. The term originated
with Herodotus in the 5th century BCE because the shape of deltas in map view can be similar to
the Greek letter delta. The shapes of deltas are subsequently modified by rivers, tides, and
waves. The three main classes of deltas are river dominated (for example, Mississippi River),
wave dominated (Nile River), and tide dominated (Ganges River). Ancient deltas contain some of
the largest and most productive petroleum systems.
Marine deposition pertains to deposition in seas or ocean waters, between the depth of low tide
and the ocean bottom.
Figure 2-3. The major compositional layers of the Earth are the core, mantle, and crust.
The lithosphere includes all the crust and the uppermost part of the mantle. The
asthenosphere forms much of the upper mantle. The Mohorovičić discontinuity
(abbreviated to Moho) is the boundary between the crust and the mantle. These
compositional and mechanical layers influence movement of the Earth's tectonic plates.
Plate tectonic theory explains such phenomena as earthquakes, volcanic or other igneous
activity, mid-oceanic ridges and the relative youth of the oceanic crust, and the formation of
sedimentary basins on the basis of their relationships to lithospheric plate boundaries. Convection
of the mantle is postulated to be the driving mechanism for the movement of the lithospheric
plates (Fig. 2-4). Global Positioning System measurements of the continents confirm the relative
motions of plates. Age determinations of the oceanic crust confirm that it is much younger than
continental crust because it has been recycled by the process of subduction, where one
lithospheric plate moves beneath another, and regenerated at mid-oceanic ridges.
At divergent margins,
• the region is initially uplifted as a result of thermal expansion (Fig. 2-5a; for example,
Colorado plateau)
• normal faults and a rift valley develop as the plates begin to separate, with volcanism and
earthquakes frequent (Fig. 2-5b, African Rift Valley)
• new oceans form as oceanic crust is produced (Fig. 2-5c, Red Sea)
• eventually a large ocean is created (Fig. 2-5d, Atlantic Ocean).
Figure 2-6. Convergent margins occur where tectonic plates move toward each other: (a)
oceanic-continental plate convergence, (b) oceanic-oceanic plate convergence, and (c)
continental-continental plate convergence.
2.5 Faults
The dynamic nature of the Earth’s crust results in the formation of faults and folds.
A fault is a break or planar surface in brittle (meaning having little or no inelastic deformation)
rock across which there is observable displacement. Depending on the relative direction of
t e n sion
ex
r e s sion
p
c om
l
latera
Figure 2-7. Fault movement depends on the force-induced stresses on the rock.
The fault block above the fault surface is called the hanging wall, whereas the block below the
fault is the footwall. The throw of a fault is the vertical displacement of the layers caused by
faulting (Fig. 2-8). The dip of a fault is the magnitude of the inclination of the fault plane from
horizontal (angle α in Fig. 2-8).
Figure 2-8. On a normal fault, the hanging wall has moved down relative to the footwall.
For a normal fault, the hanging wall moves down relative to the footwall along the dip of the fault
surface, which is steep, from 45° to 90°. Multiple normal faults can produce horst and graben
Figure 2-9. Horsts and grabens can form where the crust is being pulled apart.
A reverse fault forms when the hanging wall moves up relative to the footwall parallel to the dip of
the fault surface. A thrust fault is a reverse fault in which the fault plane has a shallow dip,
typically much less than 45°. In cases of considerable lateral movement, the fault is described as
an overthrust fault. Thrust faults can occur in areas of compression of the Earth's crust.
Movement of normal and reverse faults can also be oblique, as opposed to purely parallel to the
dip direction of the fault plane. The motion along a strike-slip fault, also known as a transcurrent
or wrench fault, is parallel to the strike of the fault surface, and the fault blocks move sideways
past each other. The fault surfaces of strike-slip faults are usually nearly vertical. A strike-slip fault
in which the block across the fault moves to the right is termed a dextral strike-slip fault. If it
moves left, the relative motion is called sinistral (Fig. 2-10). A transform fault is a type of strike-
slip fault associated with convergent and divergent plate boundaries.
Sinistral
Dextral
The presence of a fault can be detected by observing characteristics of the rocks such as
changes in lithology from one fault block to the next, breaks and offsets between strata or seismic
events, and changes in formation pressure in wells that penetrate both sides of a fault. Some fault
surfaces contain relatively coarse rubble that can act as a conduit for migrating oil or gas,
whereas the surfaces of other faults are smeared with impermeable clays or broken grains or
have had impermeable minerals deposit in the fault plane, resulting in a sealing fault that can act
as a seal.
Given the geological complexity of some faulted rocks, especially those that have undergone
more than one episode of deformation, it can be difficult to distinguish between the various types
of faults. Also, areas deformed more than once or that have undergone continual deformation can
have fault surfaces that are rotated from their original orientations, so interpretation is not
straightforward.
• Anticlines are arch-shaped folds in which rock layers are upwardly convex. The oldest
rock layers form the core of the fold, and outward from the core the rocks are
progressively younger.
• A syncline is the opposite of an anticline, with downwardly convex layers and young
rocks in the core of the fold.
Folds typically occur in anticline-syncline pairs.
The hinge is the point of maximum curvature in a fold, with the limbs on both sides of the fold
hinge. The hinge line connects the points of maximum curvature in a folded layer, and the axial
surface connects the hinge lines of the layers in a fold (Fig. 2-11). The axial surface is called the
axial plane for folds that are symmetrical and the hinge lines are coplanar.
Concentric, or parallel, folding preserves the thickness of each layer as measured perpendicular
to original bedding, whereas the layers of similar folds have the same wave shape but the
thickness changes throughout each layer, with thicker hinges and thinner limbs.
Figure 2-11. Folds are wavelike structures that typically have paired anticlines and
synclines. Axial surfaces are imaginary surfaces that connect the hinge lines along the
points of maximum curvature in a folded layer.
A dome is a type of anticline that is circular or elliptical, with the rock layers dipping away in all
directions. The upward migration of a mushroom- or plug-shaped salt diapir can form a salt dome.
Similarly, a basin is a depression, or syncline, that is circular or elliptical rather than elongate (Fig.
2-12).
Figure 2-12. Domes and basins are the circular or elliptical forms of anticlines and
synclines, respectively.
Figure 2-13. A petroleum system must have a “kitchen” where mature source rock has
reached appropriate conditions of pressure and temperature to generate hydrocarbons
that migrate into a porous and permeable reservoir rock. The reservoir rock must have a
seal to prevent the hydrocarbons from migrating through the reservoir rock. The reservoir
rock and seal must be configured such that they form a trap in which the hydrocarbons
accumulate. The timing of all these components is critical, as is the preservation of the
reservoir seal and properties over time to ensure the hydrocarbons are not lost.
Source rock is rich in organic matter which, if heated sufficiently, generates oil or gas. Typical
source rocks, usually shales or carbonates, contain about 1% organic matter and at least 0.5%
total organic carbon (TOC), although a rich source rock might have as much as 10% organic
matter. Preservation of organic matter without degradation is critical to creating a good source
rock. Rocks of marine origin tend to be oil-prone, whereas terrestrial source rocks (such as coal)
tend to be gas-prone.
Maturity refers to the state of a source rock with respect to its ability to generate oil or gas. As a
source rock begins to mature, it generates gas. As an oil-prone source rock matures, the
generation of heavy oils is succeeded by medium and light oils. Above a temperature of
approximately 100 degC [212 degF], only dry gas is generated. The maturity of a source rock
reflects the ambient pressure and temperature as well as the duration of conditions favorable for
hydrocarbon generation.
Migration is the movement of hydrocarbons from their source into reservoir rocks. The movement
of newly generated hydrocarbons out of their source rock is primary migration, also called
expulsion. The further movement of the hydrocarbons into reservoir rock in a hydrocarbon trap or
other area of accumulation is secondary migration. Migration typically occurs from a structurally
low area to a higher area because of the relative buoyancy of hydrocarbons in comparison with
A reservoir is a subsurface body of rock with sufficient porosity and permeability to store and
transmit fluids. Sedimentary rocks are the most common reservoir rocks because they have more
porosity than most igneous and metamorphic rocks and form under temperature conditions at
which hydrocarbons can be preserved.
A trap is a configuration of rocks suitable for containing hydrocarbons and sealed by a relatively
impermeable formation through which hydrocarbons will not migrate. Traps are described as
structural traps (in deformed strata such as folds and faults) or stratigraphic traps (in areas where
rock types change, such as unconformities, pinchouts, and reefs) (Figs. 2-14 and 2-15).
Figure 2-14. Structural traps are formed by the deformation of strata such as faulting or
folding.
Figure 2-15. Stratigraphic traps are formed by erosional and sedimentary processes.
Timing refers to the sequence in which the components form and processes occur (Fig. 2-16).
For example, if the source rock reaches maturity and generates oil but there is no trap formed
over the reservoir rock, then the oil will eventually seep to surface and be lost to the atmosphere
instead of accumulating.
Figure 2-16. The timing of the formation of the major elements of a petroleum system is
critical in developing a hydrocarbon accumulation. The petroleum system events can be
presented graphically to determine if the sequence may have resulted in hydrocarbon
1
accumulations. This petroleum system is in the Maracaibo basin of Venezuela.
Preservation refers to the stage of a petroleum system after hydrocarbons accumulate in a trap
and are subject to degradation, remigration, tectonism, or other unfavorable or destructive
processes.
As shown in Fig. 2-13, if the hydrocarbon volume that migrates into the reservoir exceeds the
volume that can be held, then the excess hydrocarbons will spill from the trap. The spill point is
the structurally lowest point in a hydrocarbon trap that can retain hydrocarbons. Once a trap has
1
Cross section after Parnaud, F., Gou, J., Pascual, J.-C., Capello, M.A., Truskowski, I., and
Passlacqua, H.: “Stratigraphic Synthesis of Western Venezuela,” Petroleum Basins of South
America, A.J. Tankard, R. Suárez S., and H.J. Welsink (eds.), Tulsa, Oklahoma, USA, AAPG
Memoir 62 (1995), p. 681. Stratigraphic column from Talukdar, S.C., and Marcano F.: “Petroleum
Systems of the Maracaibo Basin, Venezuela,” The Petroleum System—From Source to Trap (1st
ed.), L.B. Magoon and W.G. Dow (eds.), Tulsa, Oklahoma, USA, AAPG Memoir 60 (1994), p.
475.
2.8.1 Depth
Depth is the distance from a reference point to a target point.
The depth reference is the point from which depth is measured, at which the depth is defined as
zero. As shown in Fig. 2-17 the depth to a target horizon can be defined relative to mean sea
level (MSL), ground level (GL) if on land, or the kelly bushing (KB) of the rig used to drill a well to
the target horizon.
The depth reference for a well is typically the top of the KB or the level of the rig floor on the rig
used to drill the well. The depth measured along the well trajectory from that point is the
measured depth (MD) for the well. For Well A in Fig. 2-17, the MD to reach the target horizon is
less than for Well B even though the target horizon is at the same true vertical depth (TVD). Even
when the drilling rig has been removed, all subsequent measurements and operations in the well
are still tied in to the original depth reference. However, for multiwell studies, the depths are
usually shifted to the permanent datum, which is commonly MSL. Care must be taken in
discussing the TVD of a target because drillers generally use the KB as their reference while
geologists and geophysicists often have their maps referenced to MSL. Also note that many
horizontal wells are re-entries, where a sidetrack is drilled out of an old well. In this case the rig
that is drilling the current borehole may have a different rig floor or KB elevation compared to the
rig that drilled the original borehole. This elevation offset must be taken into account when
creating formation models and comparing MD and TVD logs.
Information regarding reference elevations such as GL above MSL, KB and drill floor above GL or
MSL is recorded on the log heading.
Horizon
Figure 2-17. Depth is the distance from a reference to a target point. Both the reference
and path to the target must be defined.
True stratigraphic thickness is the thickness of a rock layer measured perpendicular to the layer.
Although this is the most intuitive definition of thickness, determining the TST in a well requires
corrections for the dip and azimuth of both the layer and the well that intersects it.
The true vertical thickness is the thickness measured vertically at a point (Fig. 2-18). TVT is the
thickness that is measured in a vertical well through the layer.
True
vertical
thickness
Tr u
stra e
thic tigra
kne phic
ss
Figure 2-18. True stratigraphic thickness is the thickness of a rock layer measured
perpendicular to the layer, whereas true vertical thickness is the thickness measured
vertically at a point.
The TST values in an area can be plotted and contours drawn to create an isopach map (see
Section 2.9, “Contour maps.” The TVT values in an area can be plotted and contours drawn to
create an isochore map.
The measured depth thickness of a layer is the thickness of the layer measured along the
trajectory of the well that intersects that layer. Although the TST or TBT of a layer can remain
constant, the MDT of the layer intersected by a wellbore varies as a function of the dip and
azimuth of the layer as well as the dip and azimuth of the wellbore that intersects it.
Figure 2-19 shows a dipping layer intersected by three wells. Ignoring azimuthal effects in or out
of the page, Well A, which is drilled perpendicular to the layer dip, has an MDT equal to the TST.
Well B is vertical, so in this well the MDT of the layer equals the TVT. To convert the TVT to the
TST, the dip of the layer must be taken into account.
Well C, which is drilled at a low incidence angle between the well and layer, has the longest
exposure to the layer and hence the largest value of MDT.
Be
d dip
αd
eg
re
es
Figure 2-19. The measured depth thickness (solid red lines) of a layer increases with
decreasing incidence angle between the wellbore and the layer in which the measured
depth thickness is determined.
Be
dd
ip
α de
gr
ee
s
Figure 2-20. For a well drilled perpendicular to the layer, the MDT and TST are equal. A
vertical well through a layer has an MDT that equals the layer TVT at that point.
As shown in Fig. 2-20 for a well drilled perpendicular to the layer, the MDT and TST are equal:
Be
TVDT = true vertical depth thickness dd
ip
αd
eg
re
es
Figure 2-21. True vertical depth thickness is the vertical component of the MDT. It is not
the same as the TVT.
Because Well C is drilling downdip along the layer, its MDT is greater and the vertical projection
of the MDT is also greater than that in Wells A and B. A TVD log from Well C shows this layer as
considerably thicker than a TVD log from either of the other wells. To convert from TVDT to TVT,
the incidence angle and azimuth angle between the wellbore and layer must be taken into
account:
where
β = well deviation
γ = acute angle between well azimuth and dip azimuth.
Accounting for change in TVDT is particularly important when correlating between wells. If one
well has been drilled updip and another drilled downdip, then when the MD log is converted to
TVD for correlation the layer thicknesses will be different (Fig. 2-22).
From Fig. 22 it can be seen that for a well drilling updip TVDT < TVT < TST. For a well drilling
downdip in the same formation TVDT > TVT > TST. The TST remains the same in both wells but
when projected to a TVD depth scale, the layer thicknesses appear different.
Figure 2-23. Strike is the azimuth of the intersection of a plane, such as a dipping bed, with
a horizontal surface. Dip is the magnitude of the inclination of a plane from horizontal.
True, or maximum, dip is measured perpendicular to strike.
β C
o
δ
α A
ike D
Str
Figure 2-24. The true dip α is measured perpendicular to the strike; apparent dip δ is
measured at any other angle.
where
α = true dip of the layer
β = angle between the azimuth of the dipping plane and the azimuth of the apparent dip
δ = apparent dip of the layer.
AB = CD
AB = OAtan α
OC = OA/cos β.
For example, a well drilled at 45° from the azimuth of maximum (true) dip shows an apparent dip
along the well of 10°. The true dip α is derived as
The contour map most commonly used in well placement is the structure map. A structure map is
a type of subsurface map with contours representing the elevation of a particular formation,
reservoir, or geologic marker in space, such that folds, faults, and other geologic structures are
clearly displayed (Fig. 2-26). Its appearance is similar to that of a topographic map, but a
topographic map displays elevations of the Earth's surface and a structure map displays the
elevation of a particular rock layer, generally beneath the surface. Structural elevation information
is generally derived from the interpretation of seismic reflectors and formation tops identified in
well logs.
-8900
-9000
-9100
-9200
-9300
-8900
-9000
-9100
-9200
-9300
Subsurface Expression
Figure 2-25. A structure map represents the subsurface elevation of a geologic surface.
Displacement along a fault results in discontinuous contour lines. Figure 2-26 shows the structure
map and 3D representation of a simple dome structure with a planar fault that has not yet moved.
The fault is shown as a curve across the contour lines. This is because the planar fault is dipping,
so it intersects the contour lines along an arc. A vertical fault would appear as a straight line on
the structure map.
-8900
-9000
-9100
-9200
NE
LA
TP
UL
FA
-8900
0
-900
0
-910
-9200
Subsurface Expression
Figure 2-26. The structure map and 3D representation of a simple dome structure show a
fault plane, but without any movement along the fault plane.
Figure 2-27 shows the same faulted dome structure after the structure has moved along the fault
plane. In the 3D representation the contour lines remain at the same level, but the right side of
the structure has dropped. The top contour line on the right structure is now –9,000 ft, rather than
the –8,900 ft it was before the movement along the fault. However, the –9,000-ft contour can be
followed around the two sides of the dome and across the face of the fault.
-900 -9000
0
- 910 -9100
0
-9200 -9200
-9300
NE
LA
TP
UL
FA
-900 -9000
0
-910 -9100
0
0
-920 -920
0
-9300
Subsurface Expression
Structure maps often indicate where wells intersect the formation and the status of the wells
(Figs. 2-28 and 2-29).
Dry Hole
-13000
-12500
Fault@-10250
-12000
#2
-12750
#A -11500
#B
-12000
#1 0
-13250 -1300
0 Scale
-1350
1000’
Figure 2-29: A typical structure map of a small oilfield has contour lines at a 500-ft interval.
There is a fault across the middle of the structure, two producing oil wells on the crest of
the structure, and two dry holes on the flanks of the structure.
The directional driller also exploits drilling parameters such as weight on bit and rotary speed to
deflect the bit away from the axis of the existing wellbore. In some cases, such as drilling steeply
dipping formations or unpredictable deviation in conventional drilling operations, directional
drilling techniques may be employed to ensure that the hole is drilled vertically.
Although there are many directional drilling techniques, the general concept is simple: orient the
bit in the direction that one wants to drill. The most common way is through the use of a bend
near the bit in a downhole steerable mud motor. The bend points the bit in a direction different
from the axis of the wellbore when the entire drillstring is not rotating. By pumping mud through
the mud motor, the bit turns while the drillstring does not rotate, allowing the bit to drill in the
direction it points. When a particular wellbore direction is achieved, that direction may be
maintained by rotating the entire drillstring (including the bent section) so that the bit does not drill
in a single direction off the wellbore axis, but instead sweeps around and its net direction
coincides with the existing wellbore. Rotary steerable systems (RSSs) enable steering while
rotating, usually with higher rates of penetration and ultimately smoother boreholes.
3.3.1 Jetting
For jetting, the drill bit is usually equipped with small-diameter tungsten
carbide nozzles, which produce a high-velocity drilling fluid stream exiting
the bit. The jetting action cleans the bit and agitates the formation cuttings
into suspension in the mud for transport to the surface and out of the hole.
Though not as common as in the past, a bit may be fitted with asymmetric
nozzles, one large and two or more small nozzles. If drillstring rotation is
prevented during the jetting operation, the different nozzle sizes cause
greater erosion on the side where the large nozzle is than opposite the other
nozzles, enabling intentional deviation of the well (Fig. 3-9).
Figure 3-9. Jetting can be used to erode one side of the hole, resulting in a tendency for
the BHA to deflect in the direction of the “pocket.”
Figure 3-10. The wedge-shaped whipstock is used to direct the bit away from the axis of
the established borehole to drill a deviated well.
Touch Point #3
Build
Touch Point #2 Hold
Touch Point #1 Drop
Figure 3-11. The positioning and diameter of touch points on a BHA defines whether the
assembly tends to increase (build), maintain (hold), or decrease (drop) well inclination.
hold
drop
Figure 3-12. Unsupported sections of a BHA bend under the influence of gravity and
buckling loads caused by high weight on bit. The bending can result in a deflection of the
bit (shown by the dashed lines) and hence the inclination at which the well is drilled.
In a deviated hole gravitational and buckling forces bend any unsupported section of the BHA. If
two stabilizers are widely spaced with one of them relatively close to the bit, then the BHA bends
between them, resulting in the bit deflecting upward and the well inclination increasing. This is
called a build assembly (top panel of Fig. 3-12).
If stabilizers are positioned a significant distance from the bit, the length of BHA between the
bottom-most stabilizer and the bit bends slightly under gravity, resulting in a tendency for the bit
to point down and decrease the well inclination. This is called a pendulum or drop assembly
(bottom panel of Fig. 3-12).
If stabilizers are distributed relatively evenly along the length of the BHA, then there is no
tendency to either build or drop. This assembly tends to hold the well inclination and is called a
packed or hold assembly (middle panel of Fig. 3-12).
The degree to which an assembly builds, holds, or drops is called the build rate, which is the rate
of change in borehole inclination given in degrees per 100 ft [30 m]. The degree to which an
assembly turns to the right or left is called the turn rate and is also expressed in degrees per 100
ft [30 m]. The dogleg severity (DLS) of a well is the vector sum of the build and turn rates. DLS
accounts for the change in both the inclination and azimuth expressed in degrees per 100 ft [30
m].
In addition to the positioning and diameter of touch points, a number of other factors influence the
dogleg tendency of a BHA.
Larger diameter collars with thicker walls are stiffer and hence do not bend as easily as smaller
diameter or thinner walled collars (Fig. 3-13). Because of the greater flexibility of a 4¾-in
assembly compared with that of a 6¾-in or larger assembly, 4¾-in assemblies can deliver greater
doglegs, but they can also be more challenging with respect to maintaining directional control.
Power Section
Transmission Assembly
Drive Shaft
Figure 3-15. Steerable motors consist of six key elements, as shown on the cross section
(right). The longest element of the steerable motor (left) is the power section.
The power section, also known as the PDM, converts hydraulic power from the mud circulation
into mechanical power in the form of bit rotation (Fig. 3-16).
The power conversion is achieved through a progressing cavity design in which the movement of
the mud pushes on a rotor with one less lobe than the number of cavities in the stator in which it
is housed (Fig. 3-17).
Rotor
Cavities
Stator
Figure 3-17. The cross section of a motor looking down the length of the motor (left)
shows the fit of the rotor that has one less lobe than the corresponding stator has cavities,
leaving a gap for mud circulation to push on the spiraled rotor and create rotation, as
shown in the cut-away section of a motor (right).
The movement of mud from one cavity to the next turns the rotor and thus turns the bit coupled to
the rotor through the connecting rod and drive shaft. The key feature of a PDM is that the bit
rotates when there is mud circulation, even if the drillstring is stationary.
Increasing the number of lobes and cavities increases the torque available but decreases the
speed of bit rotation. Consider a unit volume of mud flowing through a motor. If the motor has a
1:2 configuration, the rotor progresses 360/2 = 180°. If the motor has a 7:8 configuration, then the
rotor progresses 360/8 = 45° for the same volume of mud. For a given mud flow rate, the
revolutions per minute (rpm) of a motor is mainly controlled by the rotor-stator configuration.
Similar to a gearbox in a car, higher gearing ratios give lower rpm and higher torque. As shown in
the bottom right panel of Fig. 3-18, torque increases and the rpm decrease with an increase in the
number of lobes and cavities in the motor configuration.
and
matching
Stators
1:2 3:4 4:5 7:8
Figure 3-18. Increasing the number of lobes and cavities increases the torque available
with a corresponding decrease in the bit rpm.
The rotor has a highly polished surface that creates a seal with the rubberized elastomer insert
forming the internal profile of the stator. The elastomer is housed in the stator tube, or “can.” The
dimensions of the rotor and stator must be carefully matched to ensure a good seal while
preventing excessive interference between the rotor and stator (Fig. 3-19).
Compression
of stator
Figure 3-19. Correct fit of the rotor to the stator is vital to motor performance. Negative
interference results in loss of power; excessive positive interference results in rapid wear
and heat generation.
Pressure differential across two adjacent cavities forces the rotor to turn, opening adjacent
cavities and allowing the fluid to progress down the length of the stator (Fig. 3-20).
FLOW
ENTRANCE CAVITIES
REGION STATOR CONTOUR
(CUTBACK)
Figure 3-20. A cross section along the length of the motor (upper) shows how the spiraled
rotor seals on the elastomer inside the stator tube so that the mud flow must turn the rotor
to gain access to the next cavity. The 3D wireframe (lower) shows how the mud-filled
cavity (red) moves along the motor as the rotor turns inside the stator.
The stage length is defined as the axial length required for one lobe to rotate 360° along its
helical path (Fig. 3-21). Stage length is also referred to as the “pitch.”
The hydraulic power extracted from the mud flow by the motor is
hydraulic power (hp) = flow rate (galUS/min) × pressure drop (psi)/1,714, (3-1)
and the mechanical power output by a motor to the bit for drilling is
for which
• Mud flow rate is controlled at surface by adjusting the rig mud pump output.
• Pressure drop is controlled by the number of stages.
• The rpm and torque are controlled by the configuration of the motor. The greater the
number of lobes and cavities, the higher the torque and the lower the rpm.
The surface-adjustable bent housing allows the angle between the bit shaft and the axis of the
motor to be changed, generally in the range from 0° to 3° (Fig. 3-22). Adjustment is performed
only at the surface, where the stator housing can be unscrewed from the adjustment ring (1 and 2
in Fig. 3-22). The ring is then lifted to disengage the alignment teeth, and the ring can then be
rotated to give the required bent housing angle (3). The adjustment ring is then slotted into
position (4) and the stator adaptor screwed back into place to lock the ring (5).
Stator adapter
Stator adapter
Adjusting
ring
Offset
Splined mandrel housing
Adjusting ring
Alignment
Offset housing teeth
Figure 3-22. The surface-adjustable bent housing allows changing the angle between the
bit shaft and the axis of the motor.
The greater the bent housing angle, the greater the dogleg that can be achieved with the motor.
However, high bent housing angles also cause significant stresses in the housing when rotary
drilling, so the bend is kept as small as possible.
By orienting the surface-adjustable bent housing in the desired drilling direction and rotating the
bit by pumping mud through the positive displacement motor, it is possible to drill in a desired
direction. The process of keeping the drillstring oriented in a desired direction while drilling is
called sliding. The orientation, or toolface, indicates the direction in which the well is to be
deviated. In deviated and horizontal wells the toolface is referenced to the high side of the hole
and is called the gravity toolface (GTF):
• 0° GTF indicates building angle.
• 90° GTF indicates turning right.
• 180° GTF indicates dropping angle.
• 270° GTF indicates turning left.
In nearly vertical holes the toolface is oriented relative to north. In this case it is called a magnetic
toolface (MTF):
• 0° MTF indicates deviating north.
• 90° MTF indicates deviating east.
• 180° MTF indicates deviating south.
• 270° MTF indicates deviating west.
Because the angle on the bent housing is small (typically less than 3°), the steerable motor can
also be rotated. This negates the effect of the bend and gives a relatively straight borehole, which
For example, if a motor slides at 5°/100 ft and hol ds angle while rotating, then to achieve a
dogleg of 2°/100 ft, the driller needs to slide for 2°/5° = 40% of the interval and rotate for the
remaining 60%. If the interval to be drilled with a 2°/100-ft dogleg is 60 ft [18 m], then the driller
slides for 40% × 60 ft = 24 ft [7 m] and rotates for the remaining 36 ft [11 m]. Depending on the
intervals involved, the driller may chose to break the slide-rotate sequence into shorter intervals
of sliding alternating with shorter intervals of rotating so that the overall change in well angle is
the same but distributed more evenly along the 60 ft than if all the sliding is performed at the
beginning.
Figure 3-23 shows a directional performance plot for a directional motor over numerous slide-
rotate sequences. The darker background shading indicates slide sequences. The GTF (pink
dots) is visible during the slide sections but not during the rotate sections because toolface is
meaningless when the bent housing is rotating. The continuous azimuth (green dots) overlies the
static survey azimuth (yellow squares connected by dark green lines). The static survey
inclination (yellow squares connected by blue lines) was taken every 90 ft [27 m]and indicates a
relatively smooth increase in inclination up to about 59°, after which the inclination is held stable.
The continuous inclination (red dots) shows the well to be far more tortuous. During the slide
sections the driller is building inclination, but the BHA drops inclination when rotating to cause the
slide-rotate sequence to result in an undulating borehole. The difference between the static
surveys and the continuous inclination and azimuth data is not unusual. The averaging caused by
the long distance between static survey stations results in a calculated trajectory that seems
smoother than is really the case.
Figure 3-23. In this directional performance plot for a directional motor over numerous
slide-rotate sequences, the oscillations in inclination (red dots) caused by the slide-rotate
sequence result in an undulating well trajectory.
It should also be noted that over sliding intervals the LWD image data is unavailable not where
the motor is sliding but over the interval where the LWD imaging sensors are sliding. This is due
to the distance between the bit and the LWD imaging sensors.
Another difficulty with slide-rotate sequences is that orienting prior to each slide section is time
consuming, and reduces the overall rate of penetration achieved for the well. In addition, because
the mud is not agitated during sliding, hole cleaning efficiency is reduced, and the probability of
becoming differentially stuck due to the pressure imbalance between the wellbore and formation
increases. Finally, the overall length of the well may be limited because static friction during
sliding, which is greater than dynamic friction while rotating, may prevent effective weight transfer
to the bit. To overcome these limitations, rotary steerable systems were developed.
A push-the-bit RSS uses pads on a bias unit to push against the borehole wall, which pushes the
bit in the opposite direction (Fig. 3-24).
Activated Pad
Figure 3-24. The three pads of a push-the-bit RSS push against the borehole wall to deflect
drilling of the well in the opposite direction.
As the system rotates, the pads must be activated in sequence to ensure consistent steering in
the desired direction. The control unit contains the electronics for control of the toolface and the
percentage of time spent steering. The system operates by diverting a small percentage of the
mud flow to activate the pads. By sensing the rotation of the BHA relative to the Earth’s magnetic
field and controlling an electric motor to rotate in the opposite direction, the control unit holds a
control valve (grey element in Fig. 3-25) geostationary. The port in the control valve is oriented
opposite the desired steering direction (toolface). Each of the three holes in the orange element in
Fig. 3-25 guides mud behind one of the three pads. A small proportion of the mud flow is thus
diverted behind each of the pads in sequence as the entry port to the piston behind each pad
rotates in front of the geostationary port in the control valve. This causes the pads to open in
sequence as they rotate into the “pushing” sector and apply force to the borehole wall opposite
the desired steering direction. The pad opens a fraction of an inch because the diameter of the
bias unit is only slightly smaller that the borehole diameter (Fig. 3-26). The drill bit, which must
have cutting elements on the side (called “side-cutting action”) as well at the front, then
preferentially cuts the rock on the side opposite the activated pad, resulting in a change in the
well trajectory.
Pad pushing
Steering
direction
Figure 3-25. The control valve (grey) is held geostationary by the control unit, which
diverts a small proportion of the mud flow behind each of the pads in sequence as each
pad’s entry port rotates in front of the hole in the control valve.
Because of the high power requirements of the control motor, the system has its own power-
generation capabilities through a mud turbine and alternator assembly. In addition to supplying
power to the motor to counter-rotate the control valve against the rotation of the collar to keep it
geostationary, the turbo-alternator system also supplies power to the electronics.
Steering is controlled by commanding the system to spend a certain percentage of the time
steering (duty cycle) in a specific direction (toolface) and the remaining time in neutral. While in
neutral the BHA could drill ahead straight if it is a hold assembly, but it could have build or drop
tendencies depending on the configuration of the three touch points. The driller needs to take the
BHA tendency into account when downlinking commands to the system.
Downlinking is performed by adjusting the mud pumps at the surface to deliver a sequence of
mud flow rate changes. The RSS detects these as changes in the rpm of the downhole turbine.
Each point on the steering command diagram, an example of which is shown in Fig. 3-27, can be
selected with a specific sequence of flow rate changes. Angles clockwise from the positive y-axis
correspond to the GTF. The distance from the center corresponds to the proportion of time spent
steering. For example, if the driller wants to build inclination as quickly as possible, the downlink
sequence commanding the RSS to operate at the point at the top of the diagram would be sent.
This corresponds to 100% duty cycle at a GTF of zero.
Downlinking to the RSS to operate at the red point in the top right quadrant would result in the
RSS spending 67% of drilling time steering up and to the right. This should result in the trajectory
building and turning to the right; however, depending on the interaction of the RSS with the
formation, the turn rates vertically and horizontally may be different. If the BHA has a strong drop
tendency, this amount of build could be required just to maintain the well at horizontal.
0.6
0.4
0.2
Build Rate 0
–0.8 –0.6 –0.4 –0.2 0.2 0.4 0.6 0.8
(fractional) –0.2
–0.4
–0.6
–0.8
Turn Rate
(fractional)
Figure 3-27. Toolface and duty cycle options are programmed into the system before it is
run in the hole. By downlinking to the system the driller can change to any other point of
the steering command diagram for the RSS. The red point is an example discussed in the
text.
The downlink command to the RSS is the means by which the system performance is adjusted.
The interaction of the entire BHA with the formation then determines the trajectory. Many systems
now incorporate an inclination hold feature. In this mode of operation the driller downlinks the
desired borehole inclination, and the RSS automatically adjusts the toolface and duty cycle
settings to maintain the requested inclination. This mode of operation can deliver very straight
trajectories and is particularly useful for well placement with deep directional measurements; for
example, the well can be drilled parallel to a remotely detected boundary.
Because of the requirement to push off the opposite side of the borehole to cause a change in the
trajectory, push-the-bit systems are sensitive to the mechanical properties of the formation.
Kicking off from vertical in an existing well can also be problematic because the pads become
unable to contact the borehole wall at the hole enlargement that occurs at the kickoff point. Point-
the-bit RSSs overcome these limitations.
A point-the-bit RSS delivers all the benefits of a push-the-bit system with reduced sensitivity to
the formation, resulting in more consistent steering, and generally higher dogleg capability. The
point-the-bit system is centered on a universal joint that transmits torque and weight on bit but
allows the axis of the bit to be offset with the axis of the system. The axis of the bit is kept offset
by a mandrel maintained in a geostationary orientation through the use of a counter-rotating
electrical motor (Fig. 3-29). Whereas the push-the-bit system keeps a control valve geostationary
to divert mud behind the pads, the point-the-bit system keeps a mandrel geostationary.
Similar to a push-the-bit RSS, high power requirements require that the system have its own
power-generation capabilities through a high-power turbine and alternator assembly. The system
also contains power electronics to control the motor and sensors that monitor the rotation of the
collar and motor (Fig. 3-28). The sensors provide input and feedback for the control of the
system.
Figure 3-29. The PowerDrive Xceed point-the-bit system uses a geostationary offset angle
between the bit and collar to create a steering tendency.
As with the push-the-bit system, the directional driller controls the dogleg by downlinking to the
system to change the proportion of steering versus neutral time.
Because there is no rotation provided downhole by either type of RSS, the entire drillstring must
be continuously rotated from surface. If additional downhole rpm is desired or surface rotation
must be kept to a minimum (such as when casing wear is a concern), a mud motor (without the
bent housing) can be used above the RSS to provide downhole rotation of the RSS assembly.
or
Figure 3-30. The relationship between the build rate and radius of well curvature can be
derived by calculating the circumference and hence the radius, R, of a vertical circle
drilled at a continuous build rate.
This simple relationship allows calculation of the well position.
Consider the problem of a well being landed at horizontal (90° inclination) just below the top of
the reservoir (Fig. 3-31).
87 degrees
2 ft Top of Reservoir
4 ft
90 degrees
R
Figure 3-31. In this example well landing problem, the well has an inclination of 87° at 2 ft
above the top of the reservoir and the plan is to land the well at 90° inclination and 4 ft
below the top of the reservoir.
The well is at 87° inclination when 2 ft [0.6 m] ab ove the top of the reservoir. The plan is to land
the well at 90°, 4 ft [1.2 m] below the top of the reservoir. If the remaining section of well is drilled
at 3°/100 ft [3°.30 m], will the well be too high or too low?
From the relationship between the build rate and radius of curvature, the radius can be
calculated:
R = 5,730/B
= 5,730/3
= 1,910 ft.
As shown in Fig. 3-32 the distance between the center of curvature and the current TVD of the
well (dashed horizontal line in Fig. 3-32) can be calculated from the right-angle triangle it forms
with the radius to the current well location and the current well TVD.
19
=
=90°-87°
3 0
73
= 5
R.cos3°
B 0
= 1910 x 0.9986
73
= 5
= 1907.4 ft
R
R
87 degrees
2 ft Top of Reservoir 2 ft R - R.cos3°
= 1910 - 1907.4 ft.
4 ft = 2.6 ft
90 degrees
Figure 3-32. Will the well land too high or too low? Simple trigonometric calculations show
that it will land too high, just 0.6 ft below the top of the reservoir.
The vertical distance between the center of curvature and the current TVD of the well is
R x cos3° = 1,910 ft x 0.9986
= 1,907.4 ft.
The TVD change from the current well location to where it will land at a continuous build rate of 3°
is R minus this distance:
In this example the TVD change from the current well location is 2.6 ft [0.8 m]. Given that the well
is currently 2 ft above the top of the reservoir, this means that the well will land only 0.6 ft [0.2 m]
below the top if the 3°/100 ft build rate is maintained. The well will land too high.
The plan is to land 4 ft below the top of the reservoir. The build rate required to achieve the
landing can be calculated by first solving for the radius, as shown in Fig. 3-33
B 0
73
= 5
R.cos3°
R = R x 0.9986
R
87 degrees
2 ft Top of Reservoir 2 ft
R - R.cos3°
= 6 ft.
4 ft
90 degrees
Figure 3-33. The vertical radius is expressed as two sections to solve for the dogleg
required to land the well in the correct position.
. The vertical radius can be expressed as two sections:
• R × cos3° above the current TVD of the well
• R – (R × cos3°) below the current TVD of the well.
The plan calls for the well to land 6 ft below the current TVD of the well, so the lower section of
the radius must equal 6 ft:
R – (R × cos3°) = 6 ft
R(1 – cos3°) = 6 ft
R = 6 ft/0.00137
= 4,378 ft
B = 5,730/R
= 5,730/4,378
= 1.23°/100 ft build rate required to land the well on target.
The following sections in this chapter review the four major capabilities of typical MWD tools:
• real-time surveys for directional control—inclination, azimuth, and toolface
• real-time power generation
• real-time mud-pulse data transmission telemetry system
• real-time drilling-related measurements—weight on bit, torque at bit, and mud pressure.
Toolface is used during the drilling of a well. Once the well has been drilled, the well inclination
and azimuth define the location of the borehole in 3D space.
Magnetic toolface is used in vertical and nearly vertical wells up to about 3° inclination to define
the azimuth relative to north in which the well deviates as drilling proceeds.
The magnetometers in the survey tool are used in conjunction with the accelerometers to
determine magnetic north and subsequently true north. The BHA is then oriented relative to true
north and the MTF transmitted in real time so the directional driller can make adjustments if
required prior to drilling.
The MTF is stated clockwise from north, so an MTF of zero indicates that the well will be drilled
due north, whereas an MTF of 90° indicates that the well will be drilled east. In the example
shown in Fig. 4-1, the MTF indicates that the well will be drilled to the southwest.
MTF
Steering
Direction
Figure 4-1. Magnetic toolface is used in vertical or nearly vertical wells to define the
azimuth relative to north of the steering direction.
The gravity toolface measurement is used in deviated and horizontal wells to define the direction
in which the borehole deviates relative to the top of the hole. The accelerometers in the survey
tool are used to define the top of the hole and the GTF is measured clockwise from the top of the
hole, looking down the hole in the direction of drilling. A GTF of zero indicates that the well will
increase inclination, whereas a GTF of 90° indicate s that the well will drill to the right. In the
example shown in Fig. 4-2, the GTF indicates that the well will both build angle and turn to the
right.
L E
F HO
O
DE
SI Steering
GH
HI Direction
GTF
Figure 4-2. Gravity toolface is used in deviated and horizontal wells to define the direction
in which the borehole deviates relative to the top of the hole.
Figure 4-3. The trajectory of the well (green line) is calculated from the survey
measurements taken at discrete points (red dots) to define the position of a well in 3D
space.
MWD depth is measured continuously by using a device that determines how far the traveling
block holding the drillpipe moves during drilling. At the end of drilling each joint, or several joints
(known as a stand) of drillpipe, the MWD depth measurement is adjusted to agree with the
driller’s depth. This generally requires applying a very small change, which is linearly distributed
back along the length of the joint or stand.
Figure 4-4. Inclination is the angle between a vertical line and the path of the wellbore at
that point.
Inclination is measured using triaxial accelerometers in the MWD tool to determine the three
components of the Earth’s gravity vector relative to the position of the tool (Fig. 4-5).
Figure 4-5. Inclination is measured using three orthogonal accelerometers in the MWD tool
to measure the Earth’s gravitational field.
G = G 2x + G 2y + G 2z ,
(4-1)
where
Gx = component of the gravity vector measured along the axis of the tool
Gy = component of the gravity vector measured perpendicular to the axis of the tool
Gz = component of the gravity vector measured perpendicular to both the axis of the tool and the
y axis of the tool.
The vector sum is compared with the expected gravitational field value at the specified location
on the face of the Earth to ensure that the three accelerometers are functioning correctly. If one of
the accelerometer readings has drifted, the vector sum does not equal the local gravity value.
To calculate the inclination, only the component along the axis of the tool (the x axis) is required.
Tg is the vector sum of the Gy and Gz components (as shown in Fig. 4-6, the components
perpendicular to the axis of the tool):
G
inclination = cos −1 x .
G
(4-2)
Gy
Gx
Tg G
Tg
Gz
Gx
Figure 4-6. The inclination is determined from the component of the gravitational vector
along the centerline axis of the tool.
When the tool is vertical the x-axis accelerometer measures the entire gravity vector so the ratio
Gx/G = 1, which gives an inclination value of zero. If the tool is horizontal then the x-axis is
perpendicular to the gravity vector and the x-accelerometer reads zero, for which the ratio Gx/G =
0 gives an inclination of 90°.
Figure 4-7. Azimuth is the angle between the north reference and a horizontal projection of
the current survey position.
The calculation of azimuth is more complicated because the Earth’s magnetic field changes
direction across the face of the Earth (Fig. 4-8). In addition, the magnetic field direction also
changes with time as the axis of the core of the Earth (magnetic north-south axis), the spinning of
which creates the magnetic field, drifts relative to the axis around which the remainder of the
planet rotates (true north-south axis).
Magnetic declination is the angle between magnetic and true north at a given point on the Earth.
Triaxial magnetometers are used to measure the components of the Earth’s magnetic field along
and perpendicular to the MWD tool. The vector sum of the three measurements equals the total
magnetic vector, H:
H = H x2 + H y2 + H z2 .
(4-3)
This value is compared with the expected magnetic field strength at that time and place on the
Earth to confirm the correct functioning of the three orthogonal magnetometers.
To calculate the well azimuth, the horizontal component of the magnetic flux line is used to define
the direction of magnetic north (Fig. 4-9):
e
Lin
lux
cF
eti
gn
Ma
Magnetic
Dip
Tangent to earth’s surface
Horizontal
Component Ea
rth
’s
Su
rfa
c e
Figure 4-9. The horizontal component of the magnetic flux line is used to define the
direction to magnetic north.
Remember that the MWD tool is measuring the three magnetic components from its position in
the deviated borehole. To be able to define the horizontal plane, the inclination of the MWD tool
must be known. The inclination data derived from the accelerometers is used to define the
inclination rotation back to horizontal. Hence, for calculations of azimuth the accelerometers and
magnetometers are used.
Although the tool measures the azimuth with respect to magnetic north, oil fields are typically
mapped against a local grid that may be slightly rotated, even from true north. There are three
possible north references (Fig. 4-10):
• true—azimuth measured with reference to true north
• magnetic—azimuth measured with reference to magnetic north
• grid—azimuth measured with reference to grid north (the top of the map, depending upon
the cartographic projection).
True North
Figure 4-10. The azimuth reference can be magnetic north, true north, or grid north.
To convert from one reference to the other, the angles between them must be taken into account:
• magnetic declination—angle D between magnetic and true north
• grid convergence—angle C between grid and true north
• Total correction = D – C.
Surveys are typically referenced to grid north instead of true north, specifically so the directional
driller can determine whether the well is on plan and within the lease boundaries.
The minimum curvature method uses the angles measured at two consecutive survey points to
describe a smooth, circular curve that represents the wellbore path (Fig. 4-11).
A1
Survey 1
I1
A2
Survey 2 I2
Figure 4-11. The minimum curvature method solves for the straightest smooth arc that can
connect two survey points while respecting the MD between them and the inclination (I)
and azimuth (A) measured at each of the survey points.
The computed wellbore curvature is usually expressed as a rate of change in the well angle (both
inclination and azimuth) per 100 ft [30 m]. Called the dogleg severity, this represents the
tortuosity of the well path.
Between each survey pair we calculate ∆x, ∆y, and ∆TVD from the MD, inclination, and azimuth
at the survey points. Summing ∆x, ∆y, and ∆TVD defines the position of the wellbore in 3D space.
Figure 4-12. The information on the first page of a survey report identifies the well,
surveyed interval, trajectory calculation methods, and references. The subsequent pages
report the measured and calculated well location.
Newer version to be inserted prior to publishing.
A typical survey report is shown in Fig. 4-12. The first page includes all the data required to
identify the well, the surveyed interval, trajectory calculation methods, and the various elevation,
gravitational, and magnetic references used. Note that the gravitational and magnetic field
strengths are in units of mGal (one thousandth of the nominal gravitational acceleration) and
HCNT (H counts, one thousandth of Earth’s nominal magnetic field strength), respectively. In the
lower right panel of the first survey page the magnetic declination and grid convergence numbers
are listed, along with the total azimuth correction applied to this survey, which is referenced to
grid north because both corrections are applied (see Section 4.3.3, “Azimuth”).
• Inclination angle—This is the well inclination. If corrections have been applied, they are
specified in the last column.
• Azimuth angle—This is the well azimuth. If corrections have been applied, they are
specified in the last column. The azimuth reference is determined by which corrections
have been applied. If only the declination correction has been applied is the azimuth
referenced to true north. If the grid convergence correction has also been applied, then
the azimuth is referenced to grid north.
• Course length—The difference in MD is between the current survey and the previous
survey. The closer the surveys are together, the more accurate the representation of the
position of the borehole. Widely spaced surveys cannot capture the undulations in the
wellbore and hence cannot deliver an accurate borehole location.
• TVD—The true vertical depth at this survey point is computed from the cumulative sum of
all the ∆TVD increments computed between the surveys to this depth. The first point in
this example identifies the survey tool type (second-to-last column) as TIP (for tie-in
point), which means that all the data for this line, including the TVD, has been entered
manually from another source.
• Vertical section—This is the projection of the straight line horizontal distance between the
origin (usually vertically below the wellhead) and the survey point onto a specified
azimuth (Fig. 4-13). The azimuth from the vertical section origin to the survey point is
specified at the bottom of the left column on the first page of the survey report.
• At azimuth—The azimuth from the origin to the survey point is the azimuth along which
the total displacement is measured.
• DLS—The dogleg severity is calculated between the present and previous survey points.
DLS indicates how sharply the well is changing direction (both inclination and azimuth).
• Survey tool type—The origin of the measured data is indicated. The first survey point in
the example in Fig. 4-12 lists TIP for tie-in point, indicating that the driller has been given
the data on the first line from which to start the trajectory calculations. This may be due to
sidetracking out of an existing well.
Le
n gth
•
al
ont
oriz
nt
e
em
True H
lac
sp
Di
ng
alo h
t ion imut
ec az
l S ve
r tica rnati
Ve alte
an
Plan View
Figure 4-13. Plan view shows displacement as a straight horizontal line between the origin
and the survey point. A vertical section is the projection of the displacement onto a
specified azimuth.
There are numerous sources of uncertainty in depth, such as drillpipe length measurement
uncertainty, pipe stretch, thermal expansion, and pressure effects. Inclination and azimuth
uncertainty may accumulate from inherent instrument accuracy limitations, instrument alignment
errors in the tool, as well as tool misalignment and sag in the borehole. Magnetic interference
from the drillstring and external sources adds further uncertainty to the measured azimuth.
A number of mathematical models can be used to estimate the uncertainty in the wellbore
position. The most recent, from the Industry Steering Committee on Wellbore Survey Accuracy
2
(ISCWSA), relies on a mathematical description of all error sources for all well types, locations,
and tool performances. The output from this mathematical model is a 3D description of the
probability distribution of the well location. Because the position of the wellbore is defined in 3D
2
Williamson, H.: “Accuracy Prediction for Directional MWD,” paper SPE 56702 presented at the
SPE Annual Technical Conference and Exhibition, Houston, Texas, USA (October 3–6, 1999).
Figure 4-14. Three-dimensional uncertainty (in this case, normal distributions in each
dimension) creates an ellipsoid of uncertainty around the computed well location.
The size of the ellipsoid of uncertainty is specified along the TVD and the semimajor and
semiminor axis (Fig. 4-15). The ellipse of uncertainty is the calculated positional uncertainty
based on the cumulative survey uncertainties at each survey station.
Wellbore
1 – TVD Radius (Vertical axis)
1 North
muth
EOU Azi
2
3
Figure 4-15. The ellipsoid of uncertainty defines the 3D volume with a specific probability
of containing the wellbore trajectory.
Tie-on Point
Survey Points
Ellipsoid of Uncertainty
Cone of Uncertainty
Figure 4-16. The ellipsoid of uncertainty at each depth includes the cumulative uncertainty
of all the previous surveys.
An alternative representation of the uncertainty is in the form of a cone of uncertainty around the
wellbore (Fig. 4-17).
Zone of Interest
Figure 4-17. The cone of uncertainty, shown in pink, defines the volume within which the
well is with a certain probability.
While the nominal well location (bright pink line inside the cone) may be in the zone of interest, if
the uncertainty on the well position is larger than the thickness of the zone of interest then there is
a certain probability that the well is not in the target because the well can be located anywhere
within the cone. This is an example of how well placement adds significant value by steering the
well according to the observed geology from the LWD responses rather than steering
geometrically based on surveys with their associated uncertainties. As the reservoir targets being
drilled become thinner and more complex, the ability to place wells based exclusively on the
geometrical projection of surveys becomes increasingly risky.
3550 3550
3500 3500
<<< S Scale = 1:50(ft) N >>>
3450 3450
D
TV
8100
3400 3400
3350 3350
3300 3300
-6800 -6750 -6700 -6650 -6600 -6550 -6500
<<< W Scale=1:50(ft) E >>>
Figure 4-18. On the plan view of a geological target (square) superimposed on a reference
coordinate grid, the planned well trajectory through the target (green line) has the ellipse
of uncertainty (the horizontal projection of the ellipsoid of uncertainty) drawn around the
planned well in the center of the target.
If the driller simply positioned the nominal location of the wellbore within the geological target, it is
possible that part of the ellipse of uncertainty would fall outside of the geological target (Fig. 4-
18). To ensure that the well is actually within the geological target the driller must reduce the
target size by the wellbore ellipse of uncertainty, resulting in a smaller driller’s target that ensures
that the well, even if on the edge of the drillers target and on the edge of the ellipse of uncertainty,
is within the geological target (Fig. 4-19).
-6800 -6750 -6700 -6650 -6600 -6550 -6500
-6800 -6750 -6700 -6650 -6600 -6550 -6500
3600 3600
3600 3600
3550 3550
3550 3550
3500 3500
3500 3500
<<< S Scale = 1:50(ft) N >>>
3450 3450
D
3450 3450
TV
D
TV
8100
81 00
3400 3400
3400 3400
3350 3350
3350 3350
3300 3300
3300 3300
-6800 -6750 -6700 -6650 -6600 -6550 -6500
-6800 -6750 -6700 -6650 -6600 -6550 -6500
<<< W Scale=1:50(ft) E >>>
<<< W Scale=1:50(ft) E >>>
Figure 4-19. Shown in plan view, the driller’s target (rectangle) within the geological target
(dashed square) ensures that the well lands within the geological target.
96 ft 96 ft
w
is d
t
lle
is
ha
as
i
Th
w
d r
s con rve
tin ously
Th
w ha w a ye
t u d
Figure 4-20. Continuous inclination and azimuth overcome the surveying phenomenon of
wellbore tortuosity that is not captured by the widely spaced static surveys.
With the increasing use of rigs equipped with topdrive drilling systems, with which a 90-ft [27-m]
stand of 3 joints of drillpipe can be drilled without having to add new drillpipe, the taking of a static
survey every 30 ft [9 m] now appears to take rig time because adding a new joint of drillpipe each
30 ft is no longer required. To maximize ROP, drillers would prefer to drill the 90-ft stand between
surveys.
The example in Fig. 4-21 demonstrates the inaccuracies that accumulate if the spacing between
static surveys is allowed to extend to 90 ft. In the early part of the well the static surveys were
taken every 30 ft, but under pressure to save time, the survey frequency dropped to 90 ft. The
nature of the minimum curvature computation causes the computed position of the wellbore to
appear deeper than that computed from the continuous survey data. The discrepancy increases
with well depth, in this case accumulating a difference of over 5 ft [1.5 m] TVD after 2,500 ft [762
m] MD. In this situation, the continuous survey representation is likely to be more accurate.
An additional concern is that because the static survey is used in the geological model to define
the 3D location of formation tops and fluid contacts, the TVD discrepancy could introduce errors
into the static and dynamic models of the reservoir.
86
cD&I
MWD
87
88
89
Inc
90
91
92
93
94
9452.00
9454.00
9456.00
9458.00
TVD
9460.00
9462.00
9464.00
9466.00
9468.00
Figure 4-21. In this example of continuous inclination, the spacing of the static survey (red
curve) in the upper panel was increased from about 30 ft to 90 ft from 11,600 ft [3,536 m]
MD. The 90-ft spacing averages out many of the inclination changes, which are captured
by the continuous Inclination measurement (blue curve). The lower panel displaying the
TVD computed from the static survey shows the well deeper than the depth computed
from continuous inclination. The difference exceeds 5 ft TVD in places.
Inclination measurements taken with survey packages located close to the bit, such as those
available with Schlumberger rotary steerable tools, significantly reduce the uncertainty on the bit
inclination allowing more accurate steering of the wellbore.
Figure 4-22. Hydraulic power from the mud flow is converted into electrical power through
use of a turbine and alternator system. The stator (blue) deflects the mud flow (from left to
right) on to the rotor (red) causing it to turn. The alternator (not visible inside the yellow
housing to the right) converts the rotation to electrical power.
There are now several methods for transmitting data from the downhole tools to surface:
• electromagnetic telemetry
• wired-drillpipe telemetry
• mud-pulse telemetry.
Figure 4-23. Electromagnetic telemetry systems induce an electric field in the earth which
is detected at surface. Data is transmitted by modulating the frequency.
Figure 4-24. Wired drillpipe uses inductive couplers at the pin (left) and box (right) of the
drillpipe to make the electrical connection between one joint of drillpipe and the next.
Mud-pulse telemetry is by far the most common real-time data transmission in use today. Mud-
pulse telemetry involves encoding data in pressure pulses that propagate up through the mud
inside the drillpipe. These pressure-pulse sequences are detected at the surface and decoded to
recreate the numerical value of the data from the downhole tools. There are three main ways of
creating a mud pressure pulse (Fig. 4-25). Positive-pulse systems impede mud flow with a poppet
valve, resulting in a temporary increase in pressure. Negative-pulse systems use a bypass valve
to bleed pressure off to the annulus, resulting in a temporary drop in pressure. Continuous wave
carrier, or siren, systems use a rotating valve assembly that alternates between opened and
closed positions, resulting in an oscillating pressure wave. Data can be encoded on the siren
system by frequency or phase modulation.
Pressure
Time
Pressure
Time
Pressure
Time
Figure 4-25. The three main mud-pulse telemetry systems are positive pulse, negative
pulse, and siren, which generates a continuous carrier wave.
Data acquired by the downhole LWD tools is compressed, encoded, and transmitted to the
surface, most commonly through a mud-pulse telemetry system. The limited bandwidth of the
current mud-pulse systems (typically 0.5 to 12 bits per second, or bps) limits the amount of data
that can be transmitted to surface in real time. Improvements continue to be made in the
telemetry rate (the number of bits that can be transmitted per second) and data compression (the
amount of data transmitted per bit), but the selection of what data is sent is still required because
bandwidth capable of transmitting all the data all the time, as is the case with wireline tools, is
unlikely to be widely available in the near future.
Real-time data is transmitted as data points (d-points), each of which represents a particular
measurement (for example, formation bulk density) or is part of a larger collection of data such as
part of an image spread across several d-points (Fig. 4-26).
The d-points are grouped into frames, which define the data to be transmitted when the BHA is in
a particular mode of operation. For example, a frame designed for use when the BHA is sliding in
a deviated well would contain the GTF and nonazimuthal formation measurements because there
is no point wasting bandwidth by sending azimuthal data, which cannot be acquired when the
BHA is not rotating. A frame designed for use when the BHA is rotating would, in contrast, most
likely contain d-points for azimuthal and perhaps image data, but it would not contain a toolface d-
point because the toolface is not of interest when the BHA is rotating.
Frame
Synchronization word
and Frame Identifier
d-points
Figure 4-26. The d-points containing the measurement data are grouped into frames.
The selection of which frame to transmit is made by the downhole tool based on the well
inclination (MTF or GTF frame), whether the tool is rotating (rotary frame), and whether the mud
flow has just started again after a period of no flow (survey and utility frames). After a few training
bits to allow the surface system to synchronize, the frame identifier is sent. Because both the
surface and downhole systems were programmed with the same frames, the subsequent stream
of bits is divided into the corresponding d-points and decoded by the surface system (Fig. 4-27).
Figure 4-27. The bit stream encoded in the mud-pulse pressure wave is divided back into
the d-points by the surface system, which was programmed with the same frame
information as the downhole tool.
Decoding involves the conversion of the binary bit stream to decimal followed by application of
the reverse of the transform that was applied to the data downhole. For example, a downhole
3 3
density measurement of RHOB = 2.4 g/cm could have 0.9 g/cm subtracted and the remainder
divided by 0.01 to give a decimal number of 150. The eight-bit binary equivalent, 10010110, is
then transmitted via the mud-pulse system.
At the surface, the binary number 10010110 is converted back to the decimal number 150 and
3
the reverse transform RHOB_RT = 0.01X + 0.9 g/cm is applied (the RT suffix identifies the
measurement as real-time data to distinguish it from the recorded-mode RHOB measurement).
3
The real-time measurement of RHOB_RT = 2.4 g/cm is then available for visualization and
interpretation. The names and descriptions of Schlumberger real-time channels are listed in
Schlumberger Curve Mnemonic Dictionary at http://www.slb.com/modules/mnemonics/.
L L
W W
D D
M M Depth measured by
W W surface sensors
D D
A
L A
L
R
W R
W
C
D C
D
Figure 4-28. A limited amount of real-time data can be transmitted to the surface through
the MWD telemetry. Recorded-mode data is retrieved from tool memory when the tools
return to the surface and includes all the data from the full suite of measurements.
When the recorded-mode data is retrieved from the downhole tool, the measurements made at a
particular time are merged with the recorded depth at the corresponding time to make the time-to-
depth conversion of the data (Fig. 4-29). As measurements are distributed along the length of the
LWD tools they will each be at a different depth at a given time, unless they are collocated as is
the case for the GR and resistivity measurements shown in Fig. 4-29.
This is an important concept for quality control of log responses. Consider a GR sensor located
10 ft [3 m] from a density sensor on the BHA. The GR and density measurements are acquired at
the same time but at different depths. If there is a problem with data transmission, the density and
GR logs will be affected at different depths, indicating that the anomalous response is time
related, not due to a formation effect. If the GR and density both show a response at the same
depth then it is highly likely that it is due to a formation feature as the data has been acquired at
the same depth by the two measurements but at different times.
Because the formation is usually uninvaded during the drilling pass, the time-lapse data is a
useful way to acquire an invaded-zone resistivity, Rxo, for use in evaluating what formation fluid
was displaced by mud filtrate invasion and how much of the formation fluid is mobile.
Drilling
Time
Figure 4-30. At any given depth there are at least two and possibly more sets of
measurements made at different times. Analyzing this time-lapse data enables the
evaluation of changes in formation properties as a function of time, which may be caused
by mud filtrate invasion or geomechanical changes to the borehole.
The higher the telemetry rate, the more data can be transferred, the faster drilling can proceed,
and the better the data sampling rate. Increased telemetry rates can be achieved by increasing
the physical number of mud-pulse bits per second or by using data compression techniques to
Figure 4-32. Increasing the telemetry rate by a factor of 4 through data compression is the
equivalent of increasing the physical bit rate by a factor of 4, either of which allows
transmission of 4 times the amount of data.
The increasing use of data compression now enables the real-time transmission of images,
waveforms, and array data, which greatly enhancing the capabilities of real-time interpretation.
Another way to express this is that there are 3.6 samples per foot for the 10-s measurement and
1.8 samples per foot for the 20-s measurement.
In many cases there are complementary measurements made downhole by the MWD tool that
assist in clarifying the behavior of the downhole system (Fig. 4-33).
x060
x080
x100
x120
x140
Figure 4-33. A drilling performance log presents downhole measurements and where
possible compares them to their surface equivalents.
The temperature of the mud in the annulus is measured at the same point as the annular
pressure. The temperature is used to correct the pressure gauge for temperature effects. The
mud temperature measured in the annulus is usually lower than the internal tool temperature
owing to heat dissipation from the electronics in the tool. The mud temperature can be
significantly lower than the static formation temperature because the mud, which is relatively cool
at the surface, is pumped through the drillstring faster than it can reach thermal equilibrium with
the surrounding formation. Hence, the mud has a cooling effect on the formation.
Friction between the drillstring and the borehole can take up some of the weight reduction seen at
the surface. By measuring the actual weight applied to the bit downhole, the driller determines
how much weight is being lost to friction.
The second track in Fig. 4-33 compares the downhole weight on bit (DWOB, blue solid curve)
and the computed SWOB (red dashed curve). The shading between them highlights any increase
in separation, which indicates an increase in friction along the borehole.
Poor weight transfer to the bit may be due to cuttings buildup in the hole, BHA diameter changes
(for example, stabilizers) hanging on ledges in the borehole, borehole collapse, or sharp doglegs,
which a relatively stiff BHA may have difficulty bending through. Particularly in high-angle and
horizontal holes, frictional effects can become so great that the BHA cannot move forward. This is
called lock-up (Fig. 4-34). Any further release of weight from the surface results in drillpipe
buckling, meaning that the well cannot be drilled any farther.
Frictional effects are greater when the drillstring is sliding than when it is rotating. This is because
rotation keeps a film of mud between the drillstring and the borehole wall which lubricates any
motion. If the drillstring is static, then there is no film of mud, so the static coefficient of friction is
higher than the dynamic coefficient of friction. This is one of the reasons that RSSs with their
continuous rotation are able to drill longer, more complicated horizontal wells than mud motors
with their slide-rotate sequences.
String in compression
High friction
No movement
Figure 4-34: Lock-up occurs when frictional forces prevent the further transfer of weight to
the bit. Any subsequent release of weight from the surface results in the drillpipe buckling.
Figure 4-35. Orthogonal magnetometers measure 90°-s hifted sinusoidal responses as they
are rotated in the Earth's magnetic field, which can be used to determine the number of
rpm of the drill collar. This type magnetometer information is also used to determine the
orientation of the LWD collar when azimuthal measurements are made.
While the average number of rpm at the surface and downhole must match, frictional effects can
cause the drillpipe to slow and even stop temporarily while turns continue to be put in at surface.
In Fig. 4-35 the surface and downhole rpm are presented in the fourth track, overlaid on the same
scale. The downhole rpm curve is smooth, indicating smooth rotation of the BHA. Erratic or
“noisy” downhole rpm is diagnostic of stick-slip.
Rpm is also a useful sliding indicator. The rpm dropping to zero but hole depth increasing
indicates that the BHA is sliding. An example of this is shown from X,053 to X,071 ft in Fig. 4-33.
4.6.4 Torque
Torque can be thought of as "rotational force," or "angular force," that causes a change in
rotational motion. Torque, τ, is given by linear force multiplied by a radius (Fig. 4-36):
τ = F × r, (4-7)
where
F = component of the force vector applied perpendicular to the moment arm
r = moment arm along which the torque is applied.
The SI unit for torque is the newton-meter (N.m). In customary oilfield units, torque is measured in
the pound-foot (lbf.ft) (also known as the "foot-pound").
In a drilling environment torque is applied to rotate drillpipe and the bit. Surface torque applied to
rotate the drillstring can be measured and plotted alongside the downhole torque measured near
the bit. An example of this is shown in the second track from the right of Fig. 4-35, where the
surface torque (red dashed curve) and downhole torque (DTOR, blue solid curve) are presented
on the same scale. The small separation between them is due to frictional losses of torque
against the wall of the borehole and around doglegs in the well trajectory. A significant increase in
the separation indicates that the surface torque is not being transferred smoothly to the bit,
possibly because of a worn bit or cuttings buildup. In extreme cases, this frictional torque can
prevent rotation of the drillpipe.
In the sliding section from X,053 to X,071 ft on Fig. 4-33, DTOR does not drop to zero when the
surface torque is no longer applied. This is because the mud motor is supplying the torque
downhole to rotate the bit. Hence, DTOR is present even though no surface torque is being
applied.
The mud pressure increases as the mud travels down inside the drillpipe according to the
equation
where
MWIN = mud weight (density) as the mud is pumped into the hole
gn = acceleration due to gravity
TVD = true vertical depth of the point where the pressure is being determined.
When the mud is flowing, the pressure is increased by the standpipe pressure applied at the
surface and decreased by frictional pressure losses along the inside of the drillstring:
internal circulating pressure = SPP + internal hydrostatic pressure – internal frictional losses. (4-9)
Additional pressure losses occur across the BHA, where components such as the MWD mud
turbine and the drilling mud motor extract power from the mud flow. There is also a significant
pressure loss across the nozzles in the bit, which accelerate the mud flow to clean the bit and
agitate the rock cuttings into the mud flow for transport back to the surface.
A strain gauge measures the mud pressure inside the tool, before the pressure loss across the
mud motor and bit occurs. As the mud flows back up the annulus between the drillstring and the
borehole wall, another strain gauge measures the annular mud pressure.
The difference between the internal and annular pressures is primarily due to the pressure loss
across the mud motor or RSS and the bit. Changes in the pressure difference can help in
diagnosing performance issues with the mud motor, RSS, and bit nozzles.
Mud returning to surface is at atmospheric, or zero gauge, pressure as it flows into the mud
tanks. The annular hydrostatic pressure at a point in the well is given by
where
MWOUT = mud weight (density) of the mud in the annulus. This is greater than MWIN as a result of
the rock cuttings that are suspended in the mud returning to the surface.
To overcome the annular frictional forces acting against the flow, the pressure at a given depth
must be increased so that the mud arrives at the surface with zero pressure:
annular circulating pressure = annular hydrostatic pressure + annular frictional losses. (4-11)
∆Pinternal friction
Figure 4-37. Annular and internal mud pressure drive the mud flow.
The equivalent circulating density (ECD) of the mud in the annulus includes the effect of cuttings
in circulation and frictional effects:
When the mud pumps are off and there is no mud flow, the frictional forces drop to zero and
some of the cuttings fall out of suspension, although some of the smaller cuttings remain in
suspension. The equivalent static density (ESD) of the mud in the annulus is given by
The driller watches the ECD to ensure that the hole is being cleaned effectively and that cuttings
are not building up against the BHA downhole, a situation that could lead to stuck pipe.
Monitoring the ECD also allows timely adjustments to the mud density to maintain pressure
control of the borehole.
Wellbore pressure control is critical for safe and smooth drilling operations. If the mud pressure is
too high the formation may fracture, resulting in the loss of borehole fluid and subsequent drilling
problems. If the mud pressure is too low then the fluid in the formation may flow into the well and
begin migrating to surface, resulting in a “kick” or loss of pressure control. Depending on the
mechanical properties of the rock, too high or too low a mud weight may result in a variety of
borehole failure modes, most of which are detrimental to efficient drilling.
Well pressure control design involves determining the upper and lower pressure limits within
which the well can be drilled safely. These limits are usually expressed as fluid densities. The
upper limit is called the fracture gradient and defines the fluid density above which the formation
fractures, resulting in mud loss. The lower gradient is usually the pore pressure gradient, below
1000 ft
Figure 4-38. Where the mud weight dropped below the lower pressure limit of the pressure
window, the deepwater well took kicks.
The overburden pressure (purple curve) is the fracture gradient and defines the upper limit of the
pressure window in Fig. 4-38. The predrill seismic estimate of the pore pressure (black curve)
defines the pressure window’s lower limit. The closeness of the two curves indicates that there is
little margin for error in mud weight. The resistivity-derived pore pressure is shown in red and the
actual mud weight profile, plotted as the annular pressure-derived ECD, is in blue. Overall, the
drilling plan succeeded in staying within the narrow pressure window. However, at two depths
where the mud weight dropped below the lower pressure limit, the well took kicks.
The right track of Fig. 4-33 displays the total flow rate (blue dashed curve), SPP (red dashed
curve), and TRPM (green dotted curve). The TRPM follows the SPP whereas the total flow is
heavily averaged.
Figure 4-39. Shock is generally quantified by peak shock (measured in gn ) and the number
of shocks measured over a specified threshold.
Vibration can be thought of as the cumulative energy in the shock or root mean square (rms) of
the drillstring response to the shock.
Shock and vibration can cause failure or damage to the BHA (collars, stabilizers, connections,
and downhole tools) and drilling bit. The potential cost impact when components in the BHA are
affected by shock and vibration is huge. Examples of these costs include
• extra rig cost from tripping a failed BHA out of the hole and running in with a new one
• twist off connections, leading to a fishing operation or lost-in-hole charge for the BHA
• overgauge hole, which increases mud and cement volumes
• inability to evaluate the reservoir as a result of poor hole quality and severely degraded
formation evaluation measurements.
Shock and vibration measurements can be used to detect both good and bad drilling practices
and trends that could lead to problems. The goal is to allow drilling to continue in the safest,
efficient manner. Detection, understanding, and mitigation of shock and vibration are a way to
achieve this goal.
Monitoring drilling mechanics, diagnosis of the cause of unwanted BHA behavior, and
subsequent mitigation of shock and vibration improve the ROP through ensuring that energy is
Figure 4-40. A normal amount of shock and vibration occurs during the drilling process as
energy is mainly used to drill rock and hence maximize the ROP.
Energy Out
Energy In
Figure 4-41. Shock and vibration in excess of what normally occurs take energy away from
the drilling process, thereby reducing the ROP.
Shock and vibration are generally caused by BHA resonances resulting from interactions with the
borehole. These can be axial, such as bit bounce; torsional, such as stick-slip; or lateral, such as
forward and backward whirl (Fig. 4-42). These complex interactions between the rotating drill
collars and the borehole wall vary with the weight on bit, rpm, lubricity of the mud, and flexibility of
the BHA, among other variables. Diagnosis of the exact cause of shock can be difficult; however,
having downhole measurements to warn of potentially destructive BHA behavior alerts the driller
that remedial action must be taken.
Figure 4-42. Various BHA resonances can cause shocks and vibrations.
As LWD measurements are made in the dynamic drilling environment they are subject to the
effects of irregular BHA motion such as shock and stick-slip. Changing mud conditions such as
mud temperature, density, and resistivity variations during drilling may also affect LWD logs.
Generally, borehole conditions are more stable during wireline acquisition.
Unlike wireline logging, where a standard tool size is run in a wide range of borehole diameters,
LWD tools must have the same diameter as the other collars in the BHA, which vary from one
hole diameter to the next. Consequently, LWD tools come in a range of sizes. The diameter in
inches of the LWD collar is typically specified as part of the name of the tool. For example, the
4¾-in [12.1-cm] diameter arcVISION Array Resistivity Compensated tool is named the
arcVISION475* tool.
Table 5-1Error! Reference source not found. shows the common tool sizes and normal
borehole diameter range in which they are operated.
Table 5-1. Common LWD Tool Sizes and Corresponding Hole Diameters
Nominal Tool Diameter (Mnemonic), in [cm] Common Hole Diameter Range, in [cm]
3 1/8 (312) [7.9] 3 3/4 to 5 7/8 [9.5 to 14.9]
4 3/4(475) [12.1] 5 3/4 to 6 ¾ [14.6 to 17.1]
6 3/4 (675) [17.1] 8 1/4 to 9 7/8 [20.9 to 25.1]
8 1/4 (825) [20.9] 10 1/2 to 12 ¼ [26.7 to 31.1]
9 (900) [22.9] 12 1/4 to 17 ½ [31.1 to 44.4]
The classic triple-combo measurements (resistivity, density, and neutron) are the minimum set of
inputs to solve for water saturation. Usually the bulk density measurement, ρb, is used to derive
the formation porosity:
where
φd = density porosity
ρb = bulk density
ρfluid = fluid density
ρmatrix = matrix density.
The neutron and density measurements are usually presented on a lithology-compatible scale
(for example, limestone-compatible scale) such that in a freshwater-filled formation of the
selected lithology, the neutron and density measurements overlie. Any separation between the
two curves is an indication that either the fluid in the pore space is not fresh water or the matrix is
not that assumed to create the overlay scale. These separations help identify when the ρfluid or
ρmatrix terms in density-porosity Eq. 5-3 need to be reviewed to obtain the correct porosity.
The photoelectric factor (PEF) and its volumetric equivalent, U, which are generally available with
conventional density measurement tools, add information to help evaluate complex lithologies so
that the correct matrix density can be entered into the density-porosity equation and hence an
accurate porosity determined.
The resistivity measurements are used to evaluate the true formation resistivity, Rt. In
consideration of the invasion of mud filtrate into the formation during drilling, modern resistivity
tools measure the formation resistivity at multiple depths of investigation to characterize and
correct for the near-wellbore invasion (Fig. 5-1). For LWD measurements close to the bit the
invasion is often minimal, allowing direct formation measurements and revealing differences with
subsequent wireline measurements.
In addition to the formation porosity, determined using a combination of the density, neutron, and
photoelectric measurements, and a formation resistivity representative of the uninvaded formation
resistivity, the Archie parameters a, m, and n need to be known (for example, from core analysis)
or assumed (the default values are 1, 2, and 2, respectively). The formation water resistivity, Rw,
at downhole conditions also needs to be known. Rw is usually calculated based on downhole
temperature and pressure and the water salinity determined from water samples.
Archie’s equation can then be solved to find the proportion of the pore space filled with water,
otherwise known as the water saturation, Sw. The remaining pore space is assumed to be filled
with hydrocarbons. Hence, the hydrocarbon saturation, Shc, is determined as
The volume of hydrocarbons per unit volume of formation, Vhc, is given by the porosity of the
formation multiplied by the hydrocarbon saturation:
Given the volume of the reservoir, usually derived from seismic interpretation along with well-to-
well log correlation, the total volume of hydrocarbons in place is the volume of the reservoir
multiplied by the volume of hydrocarbons per unit volume of formation. This hydrocarbon volume
is generally corrected for the change in volume of the hydrocarbons as they move from downhole
to surface conditions. If the hydrocarbon is oil, the volume is given as the stock-tank oil in place
(STOIP). If it is gas, the volume is quoted in standard cubic feet (scf) or standard cubic meters
(scm):
where
Vreservoir = reservoir volume
B = coefficient accounting for the change in volume that occurs when the hydrocarbons move
from downhole to surface conditions.
As outlined previously, the original objective of migrating logging measurements from wireline
tools to drill collars was to obtain the inputs for formation saturation evaluation with minimal
additional rig time, less invasion, and better borehole conditions than are generally present when
data is acquired after the formation has been open to fluid invasion and borehole degradation.
With the introduction of MWD and LWD tools in the late 1980s it became apparent that the
application of LWD measurements could take advantage of the unique real-time capabilities.
LWD measurement utilization expanded from formation evaluation to real-time structural
assessment of the well position within the geological sequence and subsequent adjustment of the
drilling trajectory to place the well in the desired location.
While the formation measurements are made downhole, depth and a number of measurements
derived from it are measured at surface. Unlike wireline acquisition, where high-telemetry
bandwidth enables all the data to be transmitted and depth stamped almost instantaneously,
LWD measurements are recorded downhole against time while depth is recorded at surface
against time and the time index used to merge the data to produce a measurement versus depth
log.
5.4.1 Depth
The measured depth of a point in a well as defined by the driller is the cumulative length of
drillpipe that has been run in the hole to position the bit at that point. To make this measurement
each piece or joint of drillpipe is manually measured at surface with a tape measure and the
specific length of each piece of pipe added to the tally as the pipe is run in the hole. No
corrections for temperature, pressure, stretch, or compression are applied.
LWD and MWD depth is measured continuously using a device that determines how far the
traveling block holding the drillpipe moves during drilling. At the end of drilling each joint, or
several joints (stand) of drillpipe, the LWD or MWD depth measurement is adjusted to agree with
the driller’s depth. This generally requires applying a very small change, which is linearly
distributed back along the length of the joint or stand.
Depth is measured against time. The downhole measurements are recorded against time using a
clock in the tool, which is synchronized with the clock at the surface. The time index is then used
to merge the measurement-time data with the time-depth data to give the measurement-depth
data that is displayed on a depth log.
Density Sensor
Measurement Depth
100 ft behind the
Formation bit
is drilled Formation is
measured by
Gamma Ray
(6 minutes later)
Gamma Ray
Sensor
10 ft behind the bit
Figure 5-2. Time-after-bit example shows that the same formation depth is measured by
the various sensors at different times.
A number of the drilling-related measurements are presented on both time and depth indices.
Shock, for example, is relevant both as time data to show the historical performance of the BHA
and the shock to which it was subjected and as depth data to show if there is a correlation
between the formation type and the shock.
On retrieval of the BHA to surface, the memory of the tools is recovered. This is called the
recorded-mode data and it consists of all the measurements. Though there can be small
differences associated with limited environmental corrections on the real-time data, the recorded-
mode data should closely match the real-time data. The recorded-mode data, with a significantly
greater number of data points, usually has better resolution than the real-time data.
Figure 5-3. An LWD log has additional information that helps in understanding the log
response.
When interpreting LWD measurements there are several pieces of additional information that
should be reviewed to help understand the log responses (Fig. 5-3).
The TVD curve should be presented on the log for correlating changes in the well trajectory to
changes in the log responses. For example, a well that is cutting up and down across the same
layer has a log response that shows several layers. The trajectory information is required to
determine whether there is only one layer or many. As the formation may not be horizontal,
correlation of updip and downdip events should not be performed based on corresponding TVD
values, but on log character that is seen to repeat as a mirror image. As the incidence angle
between the borehole and formation may be different when drilling down through a layer than
when drilling up through it, the log response may appear to be stretched (lower incidence angle)
or squeezed (higher incidence angle) when comparing the mirror response to the original
response.
The rpm indicator for the corresponding tool is required whenever an azimuthal measurement is
presented. Each azimuthal tool has its own rpm indicator. As shown in the depth track of Fig. 5-
3Figure 5-3, an rpm curve that drops to zero indicates that the corresponding azimuthal sensors
are not rotating across that interval and hence the azimuthal measurements are not available.
Whether the tool is rotating must always be checked before attempting to interpret an azimuthal
measurement. In zones where sliding occurs, the azimuthal measurement sensors may not be in
good contact with the borehole wall and hence may not give a valid formation measurements. An
unstable, noisy, or oscillating rpm curve is a good indicator of uneven BHA rotation, usually
caused by stick-slip. This behavior degrades the azimuthal measurement response, as does
severe lateral shock.
Tick marks indicate the depth at which each measurement was acquired. Each measurement has
its own tick marks. Widely spaced tick marks warn that the data may be undersampled over the
interval. Close spacing of the tick marks, as seen in Fig. 5-3Figure 5-3, indicates that there is
sufficient data sampling to give a representative measurement of the formation.
Time after bit (TAB) indicates the formation exposure time prior to the measurement being taken
at that depth, as outlined in Section 5.4.3, “Time after bit.”
Shock should be presented to enable the interpreter to assess whether the BHA was moving
smoothly during acquisition. Severe shock degrades measurement response and is likely to lead
to downhole tool failure.
In addition to these general log quality control (LQC) indicators, there are tool-specific quality
control guidelines, which are listed in each tool’s Log Quality Control Reference Manual.
The initial measurements were GR and simple resistivity curves, which were used more for
correlation than formation evaluation. Gradually, sophisticated array resistivity, density, and
neutron tools have been added to the measurement portfolio, making it possible to solve Archie’s
equation for formation fluid saturations using LWD data alone. Acoustic measurements such as
LWD sonic and seismic services have been added, as have elemental spectroscopy, sigma
(thermal neutron capture cross section), and formation pressure measurements, to further
expand the use of LWD data.
The information presented to the well placement team can at times be overwhelming, so it is
important to prioritize and compartmentalize the wellbore information into well placement
information and formation evaluation information. Well placement decisions often require a quick
response, and trying to wade through data that does not truly affect the steering decision
interferes with the decision-making process. Reservoir evaluation can wait until after the steering
decision is made.
Another important consideration is that LWD tools provide raw data from the wellbore, and this
data must be interpreted before a decision can be made. Interpreting horizontal data can be
difficult at times, so it is imperative that the proper evaluation methods be used.
The GR measurement counts the number of gamma rays emitted from the disintegration of the
three naturally occurring radioactive isotopes commonly found in Earth formations: thorium (Th),
uranium (U), and potassium (K).
The GR can also be used to determine the proportion of clay in a formation because there is
usually a higher concentration of these radioactive elements in clay. However, some uranium
salts are soluble in water and can migrate through the formation dissolved in water before being
deposited elsewhere. Zones enriched with uranium by this process have a higher GR reading
than their clay content would normally have and hence may be mistaken as having a higher clay
content than is actually the case. This situation can be resolved by using a spectral GR
measurement, which distinguishes the energies of the incoming gamma rays and determines
whether the gamma rays originated from a uranium, thorium, or potassium atom. The relative
proportions of thorium and potassium can then be determined and the contribution from uranium
removed, resulting in a more robust clay indicator.
The GR is a useful measurement for well placement for the following reasons:
• GR is relatively unaffected by fluid saturation and porosity variation. As a result, it tends
to have consistent character throughout the reservoir.
• GR responds with minimal shoulder bed effect because it has a shallow depth of
investigation.
• GR can resolve much finer layering in a horizontal well than can be seen on a GR log
from a vertical well. There are two reasons for the improved measurement response in
horizontal and high-angle wells:
o The horizontal wellbore is at a low incidence angle to the beds. For example, a 6-
in– [15-cm-] thick bed is traversed by a 8.5-inch [21.6 cm] wellbore over 69.2 ft
[21.1 m] of measured depth when the well is within 1° of formation dip (Fig. 5-4),
which essentially increases its visibility to the tool.
830.8 inches
= 69.2 ft
487 inches
= 40.6 ft
8.5-inch bo 343.8 inch
rehole es = 28.6 ft
Figure 5-4. A 6-in- [15-cm-] thick bed is traversed over a measured depth of almost 831 in
[2110 cm] if an 8.5-in. [21.6 cm] well intersects the layer at a 1° angle.
LWD resistivity tools are grouped in two categories: laterolog and propagation.
Laterolog resistivity measurements are made by pushing electrical current from an electrode,
across the borehole, and into the formation, with the current returning to an electrode on the tool.
This laterolog current path measures the formation resistivities in series (Fig. 5-5). Hence
laterolog measurements are suitable for logging in conductive muds, highly resistive formations,
and conductive invasion. If the resistivity of the invaded zone, Rxo, is greater than Rt, the laterolog
measurements will be very sensitive to the highly resistive invaded zone and not be as sensitive
to the lower resistivity uninvaded formation.
Propagation resistivity measurements are derived from the changes to the phase and amplitude
of an electromagnetic wave as it propagates through a formation. The propagating wave induces
current loops in the formation that circle the body of the tool. These propagation-induced currents
measure the formation resistivities in parallel (Fig. 5-5). Electromagnetic propagation tools work
best in highly conductive formations and can operate in both conductive and nonconductive
muds, provided that the contrast between the mud and formation resistivities is not so high that
the induced currents tend to set up in the borehole.
Borehole
Invaded Formation
Uninvaded Formation
Rt
R xo
olog
Rm Later nse
e sp o
R
Rm
Rxo
Rt e
pons
o n Res
agati
Prop
Figure 5-5. Laterolog currents measure the formation resistivities in series whereas
propagation currents measure the formation resistivities in parallel.
The resistance measured by the tool is given by the voltage drop between the return and source
electrodes divided by the source current:
V
r= , (5-7)
I
where
r = resistance measured by the tool (ohm)
V = voltage drop between the source and return electrodes (volt)
I = current flowing from the source to the return electrode (ampere).
The resistivity (expressed in ohm-m) of the formation is a property of the material, whereas the
resistance (expressed in ohm) measured by the tool also depends on the volume measured. The
two are related by a system constant, called the k-factor, which in simple cases is the length
between the measurement electrodes divided by the area that the current passes through:
L
r=R , (5-8)
A
where
Electrical current always follows the path of least resistance. In a homogeneous formation the
current is evenly distributed around the tool, but in layered formations current “squeezes” into
conductive beds, distorting the electric field. Resistive beds have the opposite effect: The current
avoids them and preferentially flows in the more conductive beds. These effects, called squeeze
and antisqueeze, respectively, must be taken into account in interpreting laterolog responses
(Fig. 5-6).
Figure 5-6. Squeeze (left) and antisqueeze (right) effects result when the measurement
current (red lines) follows the path of least resistance through the more conductive layer
or fracture.
Electric currents can be tightly focused, making them suitable for formation imaging.
Consequently LWD laterolog tools generally provide resistivity images around the borehole as
well as quantitative formation resistivity data. Resistivity images are particularly useful for the
identification of conductive features such as open faults and fractures because of the squeeze
effect. Even fractures smaller than the imaging button size can be identified, because the
measurement current takes the path of least resistance along these conductive mud-filled
features. Having images from multiple depths of investigation allows evaluation of whether a
feature observed on a shallow image continues deeper into the formation. Drilling-induced
features such as borehole breakout can be distinguished from original formation features through
the comparison of images from multiple depths of investigation.
The focusing of laterolog measurements also enables evaluating the formation resistivity at
various azimuths around the borehole. Measurements from a traverse at a low incidence angle
through layers can be used to assess the resistivities of the layers above and below the wellbore
independently, yielding better layer definition and reserves estimates than if an average
resistivity, which mixes the layer responses, is used.
The GVR* resistivity sub of the GeoVision* imaging-while-drilling tool acquires five independent
laterolog resistivities simultaneously:
• bit resistivity—the bit is used as a measurement electrode
• ring resistivity—a cylindrical electrode provides a focused laterolog resistivity
• button resistivities at three depths of investigation—azimuthally focused electrodes
provide azimuthal resistivities and images.
The bit resistivity measurement is useful for early detection of significant resistivity changes in the
formation. Figure 5-8 shows an example of where the resistivity-at–the-bit (RAB) measurement
was used to detect the resistivity change as the well penetrated the top of a high-resistivity
reservoir layer from a low-resistivity shale. Drilling was stopped after only 9 in [23 cm] of reservoir
penetration, and coring was able to proceed from very close to the top of the reservoir.
As the laterolog tool pushes current toward the bit, any conductive material below the tool, such
as the housing of a motor or RSS, becomes part of the measurement electrode. Lengthening the
measurement electrode degrades the vertical (axial) resolution of the measurement and moves
the measure point back from the bit to the midpoint between the bit and lower laterolog tool
toroid. When this occurs the bit resistivity should not be used for well placement purposes
because contact of the long measurement electrode with several different resistivities
simultaneously can result in complex responses. If the bit resistivity is to be used for accurate
geostopping, the laterolog tool should be placed as close to the bit as possible.
The ring and button resistivities can be used only in conductive mud because they require a
conductive path to the formation and back. Having azimuthally focused measurements with at
least three different depths of investigation allows a simple piston invasion profile, such as that
shown in Fig. 5-1, to be solved in each of the sectors around the borehole. The three unknowns
of Rt, Rxo, and the radius of invasion, ri, are solved using the three button measurements. In this
way an invasion-corrected value of Rt can be determined in each of the sectors around the
borehole for use in formation evaluation.
X00
m
X10
m
Figure 5-9. A typical LWD laterolog presentation shows the GR and ROP (Track 1);
azimuthal average button, ring, and bit resistivities (Track 2); tool rpm (depth track); true
dips of features picked from the images (tadpoles in Track 3); and shallow, medium, and
deep button resistivities and azimuthal GR images (Tracks 4 through 7).
The complex conductance, Z, is given by the vector sum of the conductance and capacitance:
Z = G + iX , (5-11)
and the total current, IT, that flows through a formation is given by
A
I T = [σ + iωε ] V . (5-12)
L
where
V = oscillating voltage applied to the formation.
ε r.
Ωm)
(Ω
Figure 5-11. The Schlumberger correlation between resistivity and the relative dielectric
constant, εr, was derived from hundreds of core samples.
There are two approaches to deriving formation resistivity from the behavior of an
electromagnetic wave passing through rock:
• Induction measurements use the difference in the magnetic field between two receivers
that is caused by eddy currents induced in the formation.
• Propagation measurements measure amplitude and phase-shift differences between the
receivers.
If a drill collar was used to employ the same method then a similar precision for coil placement
would be required. In the harsh drilling environment, a drill collar striking the wall of the borehole
can easily produce 100-g shock, which is more than enough to ruin any precise coil positioning.
LWD tools must use a scheme in which positional stability is not as demanding. This is
accomplished by using a simple transmitter and receiver-pair arrangement. Precise coil
placement does not matter because the phase shift and attenuation are measurable with a simple
pair of coils: both quantities increase rapidly with frequency. The 2-MHz frequency has been
selected as a compromise between minimizing frequency effects (dielectric) on the measured
The propagation measurement is performed using loops of wire around a collar to transmit an
electromagnetic wave into a formation. The difference in the phase (phase shift) and amplitude
(attenuation) across a pair of coil receivers is measured (Fig. 5-12). The phase shift and
attenuation are related to the formation resistivity.
Figure 5-12. The phase shift and attenuation of an electromagnetic wave propagating from
a transmitter through a formation are measured at a pair of receivers.
Because of this decrease, attenuation resistivities should not be used above approximately 50
ohm.m.
Figure 5-15. Applying MBHC (bottom) cancels out artifacts in propagation phase-shift
resistivity (Rps) logs.
The first three factors are controlled by tool design. The names of the Schlumberger propagation
resistivity measurements encapsulate this information. For example, the resistivity measurement
labeled P34H refers to the phase-shift (P) resistivity measured with 34-in transmitter-receiver
spacing at high (H) frequency (2 MHz). An attenuation (A) resistivity measurement made with 22-
in transmitter-receiver spacing at 400 kHz (L) would be labeled A22L.
Despite being measured on the same electromagnetic wave, the phase and attenuation
measurements have independent depths of investigation. Lines of equal phase are spherical in
nature because the wave travels with equal speed in all directions (Fig. 5-16, left). The
corresponding phase-shift resistivity measurement is relatively shallow and axially focused.
However, lines of equal amplitude form a toroidal shape around the transmitter because the
amplitude is related to the energy of the wave and the tools are designed to deliver maximum
energy in the radial direction. The attenuation measurement is relatively deep but less axially
focused (Fig. 5-17).
Phase Attenuation
Figure 5-17. The phase and attenuation measurements provide two independent volumes
of investigation. The phase measurement is shallow and axially focused, and the
attenuation is deeper but less axially focused.
For the range of transmitter-receiver spacings in common use, all the attenuation measurements
are deeper than the phase-shift measurements (Fig. 5-18).
Figure 5-18. The radius of investigation of both phase-shift and attenuation measurements
increases with increasing formation resistivity. This plot is for a 6.75-in [17.15-cm] tool at 2
MHz.
For propagation measurements the radius of investigation is defined as the distance from the
borehole at which 50% of the measurement response comes from closer to the borehole and
50% comes from deeper into the formation. Figure 5-19 shows the increased radius of
investigation obtained when operating at 400 kHz.
Figure 5-19. The 400-kHz radius of investigation is deeper than that of the corresponding
2-MHz measurements in Fig. 5-18. This plot is for a 6.75-in [17.15-cm] tool at 400 kHz.
The depth of investigation of both the 2-MHz and 400-kHz measurements increases with
increasing formation resistivity (Figs. 5-18 and 5-19). The measurement current induced in the
formation by the transmitter seeks the path of least resistance. At low resistivities the current
remains relatively close to the tool. As formation resistivity increases, the current spreads over a
larger area. As outlined in Section 5.9.1, “Laterolog resistivity,” the resistance of a rock is given by
Eq. 5-8.
For high-resistivity formations, the current spreads out over a larger area to reduce the total
resistance of the path that it traverses around the tool. This spreading of the induced current
results in deeper measurements with increasing formation resistivity as the current spreads
deeper into the formation. It also results in reduced axial resolution with increasing formation
resistivity as the current spreads along the tool.
In summary:
• Attenuation measurements are deeper than phase measurements.
• Depth of investigation increases with increasing transmitter-receiver spacing.
• Depth of investigation increases with decreasing transmitter frequency.
• Depth of investigation increases with increasing formation resistivity.
The axial resolution of a propagation resistivity measurement is controlled by four main factors:
• phase or attenuation measurement of the wave
• receiver-receiver spacing
• transmitted wave frequency
• formation resistivity.
Figure 5-20 shows the axial response function of typical phase-shift and attenuation resistivity
measurements. The axial response function can be thought of as the window along the length of
the tool through which the measurement of the formation is made. The sharper the response
function, the thinner the formation layer that the measurement can uniquely resolve. If a layer is
thinner than the axial window, then it is averaged with the layers above and below it. The phase-
shift resistivity response (top) has a thinner window (better resolution) than the attenuation
measurement (bottom). This resolution corresponds with the volume of the response for the two
measurements (Fig. 5-17).
There are several different definitions of axial resolution (Fig. 5-20). First, it is the interval within
which a large percentage, typically 90%, of the axial response occurs (quantitative resolution).
Second and most commonly quoted numerically, it is the width at the 50% point of the axial
response function (width at half maximum). Third, it can refer to the smallest bed thickness for
which a significant change can be detected by the measurement (qualitative resolution).
Attenuation
Resistivity
Axial
Response
Function
Depth (inches)
Figure 5-20. The axial resolution of the phase shift-measurement (top) is sharper than that
of the attenuation measurement (bottom).
Table 5-2 lists the axial resolution, defined as the width of the axial response function at half
maximum, for various phase and attenuation measurements in typical formation resistivities. As
shown graphically in Fig. 5-20, phase-shift resistivities have a sharper axial resolution than that of
attenuation resistivities.
The axial resolution changes only slightly with varying transmitter-receiver spacing. This is
because the measurement is taken between the pair of receivers. The axial resolution is strongly
dependent on the spacing of the two receivers, but because this is fixed at 6 in [15 cm] for most
propagation tools, this is not a factor that needs to be considered for interpretation. The distance
to the transmitter has only minimal influence on the axial resolution.
At comparable conditions, the 2-MHz measurements have sharper axial resolution than the 400-
kHz measurements. In general, the 2-MHz measurements are sharper and shallower than their
400-kHz counterparts.
The axial resolution degrades with increasing formation resistivity. As outlined in Section 5.9.2.4,
“Depth of investigation,” this is because the current induced in the formation spreads out to
reduce the total resistance it experiences as it circulates in the formation around the tool.
R = 1 ohm.m
Phase-shift resistivity 2 MHz 0.7 [0.2] 0.7 [0.2] 0.7 [0.2] 0.7 [0.2] 0.7 [0.2]
Attenuation resistivity 2 MHz 1.8 [0.5] 1.8 [0.5] 1.8 [0.5] 1.8 [0.5] 1.8 [0.5]
Attenuation resistivity 400 kHz 3 [0.9] 3.5 [1.1] 4 [1.2] 4 [1.2] 4 [1.2]
R = 10 ohm.m
Figure 5-21 shows the 2-MHz propagation resistivity response to a 4-ft [1.2-m] layer of 100
ohm.m and another of 1 ohm.m sandwiched between 10-ohm.m layers. The phase-shift
resistivities with their better axial resolution get closer to Rt in the thin beds. The shorter
transmitter-receiver spacing of the 16-in [41-cm] phase-shift resistivity gets closer to Rt than the
40-in [102-cm] phase-shift resistivity because of its slightly better axial resolution. The attenuation
resistivities, which have less axial resolution, read considerably lower than Rt in the 100-ohm.m
layer. The axial resolution window for the attenuation resistivity includes both the 100-ohm.m
layer and the 10-ohm.m layers on either side. The current induced by the tool preferentially flows
in the lower resistivity 10-ohm.m layers on either side of the 100-ohm.m layer, resulting in the
lower resistivity reading.
Both the phase-shift and attenuation measurements get close to Rt in the 1-ohm.m layer because
it is relatively conductive compared with the surrounding 10-ohm.m layers. The measurement
current, seeking the path of least resistance, preferentially flows in the low-resistivity layer.
From this example it can be seen that axial resolution effects and the path of least resistance for
the induced currents within the axial resolution window must be considered in interpreting
resistivity around thin beds.
In summary:
• Phase measurements have sharper axial resolution than the corresponding attenuation
measurements.
• Axial resolution sharpens with increasing transmitter frequency.
• Axial resolution sharpens with decreasing formation resistivity.
• Axial resolution sharpens with closer receiver-receiver spacing (this is fixed for a tool).
• Axial resolution sharpens slightly with closer transmitter-receiver spacing.
The effect can increase or decrease the apparent resistivity response. Figure 5-22 shows the
borehole corrections for the phase-shift resistivities of the EcoScope* multifunction logging-while-
drilling service as a function of varying borehole diameter, mud resistivity, and formation
resistivity. However, the corrections are not always in the same direction. The magnitude of the
correction increases with increasing contrast between the mud and formation resistivities.
Rt / PxxH
9”
Rt / PxxH
10”
Rt / PxxH
Figure 5-22. Borehole corrections for phase-shift resistivities from a 6.75-in propagation
resistivity array with varying borehole size, mud resistivity, and formation resistivity.
Measurements denoted with the _UNC qualifier (for example, P40H_UNC) have been borehole
compensated (refer to Section 5.9.2.3, “Borehole compensation”) but not borehole corrected.
Borehole-corrected curves do not have the qualifier (for example, P40H).
In summary:
• A conductive borehole acts as an alternative path for propagation measurement currents.
Borehole effect may increase or decrease the apparent resistivity response.
• Borehole-compensated but uncorrected resistivities are denoted with the _UNC qualifier.
• Borehole-corrected resistivities do not have the qualifier (e.g., P40H).
In this case the 400-kHz resistivities must be used for formation resistivity evaluation.
2 MHz
Frequency Used
400 kHz
Figure 5-24. Blended resistivities use the 400-kHz resistivities below 1 ohm.m and the 2-
MHz resistivities above 2 ohm.m. Between these blending thresholds, a linear combination
of the resistivities is used.
The blended resistivities curves (Fig. 5-25) are labeled in the same way as the single-frequency
curves except that the final letter is B (for blended) rather than H (high for 2 MHz) or L (low for
400 kHz). For example, the P40H and P40L resistivities would be blended into a single resistivity
labeled P40B.
Evaluation of the invasion profile is one of the primary reasons for the development of multiple
depths of investigation for resistivity tools. A greater proportion of the shallow measurement
response comes from the invaded zone than for deeper measurements. Hence conductive
invasion causes a spread of resistivities, with the shallow measurements reading lower
resistivities than the deep measurements. This creates a conductive-invasion resistivity profile
(left side of Fig. 5-26). If the invasion is very local around the borehole, the separation on the
phase-shift resistivities may be significant but the separation on the deeper attenuation
measurements relatively small. Deep invasion results in significant separation in both the phase
and attenuation measurements.
Resistive invasion is characterized by the reverse order of the resistivities. In resistive invasion
the shallowest resistivity reads highest and the deepest resistivity reads lowest (right side of Fig.
5-26).
Rphase
Invasion - Conductive Invasion - Resistive
Rattenuation
Rattenuation
Depth (m) Depth (m)
Figure 5-26. A conductive invasion profile (left) has the shallowest resistivity reading
lowest and the deepest resistivity reading highest. A resistive invasion profile (right) has
the opposite order, with the shallowest resistivity reading highest and the deepest
resistivity reading lowest.
Multiple resistivities are required to solve an invasion model so that Rt can be determined. The
simplest invasion model is the piston invasion profile (left side of Fig. 5-27). By using resistivities
that have been corrected for borehole effects (using the known mud resistivity, Rm, and hole
radius, rh), the three unknowns Rt, Rxo, and ri can be solved with the three resistivity
measurement inputs.
The GVR resistivity sub, with three focused button resistivities, can solve a piston invasion model
in each quadrant around the borehole. This enables detecting different invasion radii and
determining Rt in each of the quadrants.
Propagation resistivity tools induce current loops that circulate in the formation around the tool.
Hence the resistivity they provide is an average from around the borehole. When solving for a
piston invasion profile with a propagation tool, it is assumed that ri, Rxo, and Rt are the same in all
directions around the borehole (axisymmetric).
The availability of more resistivity measurements enables solving more complex invasion profile
models. The ramp profile has four variables (Rt, Rxo, ri1, and ri2) and thus requires four resistivity
measurements to solve (right side of Fig. 5-27).
Resistivity, R
Rt
Resistivity, R
Rxo Rxo
Rm Rm
rh
rh radius, r radius, r
ri1
ri ri2
Rxo Rt Rxo Rt
Figure 5-27. The piston or step invasion profile (left) is the simplest invasion model and
can be determined with three resistivity inputs to solve for the three unknowns (Rt, Rxo,
and ri). Solving for the more sophisticated ramp invasion profile (right) requires four
resistivity inputs to solve for the four unknowns (Rt, Rxo, ri1, and ri2). Both examples show
conductive invasion. The approach is also valid for resistive invasion.
The piston invasion profile is the most common model used to interpret resistivity separations.
Figure 5-28 shows modeled responses for three phase-shift and three attenuation resistivities in a
piston invasion profile. A uniform 4-in [10-cm] invaded zone around the borehole with a resistivity
(green line) of Rxo = 1 ohm.m is modeled in a formation where the uninvaded formation resistivity
(pink line) varies from Rt = 100 ohm.m down to 0.3 ohm.m. Where Rt is greater than 1 ohm.m
creates a conductive invasion profile. Where Rt is below 1 ohm.m creates a resistive invasion
profile.
In the conductive invasion profile the shallower phase-shift and attenuation resistivities read lower
than the deeper measurements because the invasion has greater effect on them. In the resistive
profile the reverse separation is observed. Although ri remains constant the separation of the
resistivities increases as the contrast between Rxo and Rt increases.
In summary:
• Conductive invasion—shallow phase-shift resistivities decrease. If the invasion is deep or
the resistivity contrast high, the shallow attenuation resistivities can also decrease.
• Resistive invasion—shallow phase-shift resistivities increase. If the invasion is deep or
the resistivity contrast high, the shallow attenuation resistivities can also increase.
The discussion of axial resolution effects in Section 5.9.2.5, “Axial resolution,” considered the
effect of thin layers with differing resistivity on the propagation resistivity response when the tool
is perpendicular to the layers. Boundary effects occur when the tool and layering are not
perpendicular, so the induced measurement currents circulating around the tool are forced to
cross the resistivity contrast rather than run parallel to it (Fig. 5-29). Proximity effects occur where
the tool is not crossing the resistivity boundary but the deeper measurement currents are
responding to both the local layer (in which the tool is located) and the proximate layer (the
nearby layer affecting the deeper measurements).
Bed
boundary
effect Proximity
effect
Figure 5-29. Axial resolution, bed boundary, and proximity effects are related. Axial
resolution effects are due to averaging more than one layer within the axial window of the
measurement. Proximity effects are due to averaging of more than one layer within the
radial depth of investigation of the measurements. Bed boundary effects are combination
of axial and radial effects as the volumes of investigation of the propagation
measurements cross a change in resistivity.
When the propagation tool and formation layering are perpendicular, the induced currents
distribute themselves across the layers in inverse proportion to the resistivity. In other words, they
respond to the parallel resistivity of the two layers (Fig. 5-30):
1 Vupper Vlower
= + ,
Rmeasurement Rupper Rlower
(5-13)
where
Rmeasurement = apparent resistivity measured by the tool
Vupper and Vlower = volumetric influence of the upper and lower layers, respectively
Rupper and Rlower = resistivity of the upper and lower layers, respectively.
1
⇒ R measurement = = 1.96 ohm − m
0. 5 0. 5
+
1 50
Figure 5-30. Measurement currents parallel to formation resistivity boundaries result in a
resistivity response that is the volumetric parallel sum of the layer resistivities.
When a propagation resistivity tool approaches a change in formation resistivity with a low
incidence angle between the tool and layering, the measurement currents are forced to cross
both layers, so the resistivity response is the series sum of the resistivities (Fig. 5-31):
Rmeasurement = (Vupper × Rupper ) + (Vlower × Rlower ).
(5-14)
Figure 5-31. Measurement currents forced to cross formation resistivity boundaries result
in a resistivity response that is the series sum of the layer resistivities.
There is a significant difference in the resistivity response to the same layers as a function of the
incidence angle between the borehole and layering.
∆V = I.R1
1 ohm-m 50 ohm-m
Figure 5-32. Polarization horns result from charge buildup at the resistivity interface.
The magnitude of the polarization horn depends on the resistivity contrast between the layers and
the incidence angle between the tool and layers. In horizontal layers, the polarization horn
magnitude increases to a maximum as the well inclination approaches 90° (parallel to the layers)
(Fig. 5-33).
Figure 5-33. The polarization effect on phase-shift (upper left) and attenuation (lower left)
resistivities increases as the tool and layering come closer to being parallel. The angles
shown are wellbore inclinations to horizontal layers. The polarization effect is more
pronounced on phase-shift resistivity than on attenuation resistivity (right).
In summary:
• Bed boundary effect results in resistivity curve separation as the measurement current
crosses more than one layer.
• Polarization horns can form if
o The incidence angle between the tool and layering forces the current to cross a
resistivity contrast (approximately 45°).
o The resistivity change is sufficient to cause significant charge buildup on the
interface.
In the simple bimodal layer model shown in Fig. 5-34, the horizontal resistivity, Rh, is the
volumetric parallel sum of the layer resistivities:
1 V V
= 1 + 2.
Rh R1 R2
(5-15)
The vertical resistivity, Rv, is the volumetric series sum:
Rv = (V1 × R1 ) + (V2 × R2 ).
(5-16)
The terms “horizontal” and “vertical” resistivity are somewhat misleading. The quantities of
interest are the resistivities measured parallel and perpendicular to the layering. The horizontal
and parallel terminology is based on the assumption that the layering is horizontal, which may not
be the case. To be more accurate, Rh should be called Rparallel and Rv should be called Rperpendicular.
1 V V
= 1 + 2
R h R1 R 2
1
50 Ω-m ⇒ Rh =
25% (0.25 + 0.25) (0.25 + 0.25)
+
25% 1 Ω-m
1 50
= 1.96 ohm − m
25% 50 Ω-m
25% 1 Ω-m R v = V1 .R 1 + V2 .R 2
⇒ R v = (0.25 + 0.25) * 1 + (0.25 + 0.25) * 50
= 25.5 ohm − m
Figure 5-34. Resistivity anisotropy can be caused by alternating formation layers with
different resistivities.
10
Ω-m)
Resistivity (Ω
1
90 80 70 60 50 40 30 20 10 0
Figure 5-36. The resistivity log (left) shows the anisotropy signature of the phase-shift
resistivities increasing from shallowest to deepest and reading higher than the attenuation
resistivity . Rh and Rv were computed from the data. As with polarization horns, sensitivity
to anisotropy increases as the incidence angle between the borehole and layering
decreases (top right). TR = transmitter-receiver spacing.
Rattenuation
Depth (m)
Figure 5-37. Resistivity anisotropy causes the phase-shift resistivities to increase from
shallowest to deepest and read higher than the attenuation resistivities.
In summary:
• The tool must be at a sufficiently low incidence angle to the layering (approximately 45°)
to force induced currents to cut through multiple layers and thus exhibit sensitivity to the
anisotropy.
• Anisotropy results in a log response similar to a polarization horn extended along the
borehole (Fig. 5-36).
• Both phase and attenuation resistivities increase from shallowest to deepest, but the
phase-shift resistivities read higher than the corresponding attenuation resistivities (Fig.
5-37).
In the example shown in Fig. 5-38, the well crosses from a 100-ohm.m layer into a 1-ohm.m
layer. The phase-shift resistivities are unaffected by the proximity of the low-resistivity layer until
the tool is very close to the layer. The deeper attenuation measurements begin decreasing from
the local layer resistivity of 100 ohm.m while the lower layer is still 10 ft [3 m] TVD away. As the
well approaches the 1-ohm.m layer, the attenuation resistivities decrease further as more of their
volume of investigation is within the low-resistivity layer. The high resistivity contrast and low
incidence angle between the borehole and layering produce a large polarization horn.
As the well enters the 1-ohm.m layer, the phase-shift resistivities rapidly drop to 1 ohm.m. The
attenuation resistivities remain higher because these deeper measurements continue to have a
proportion of their response from the 100-ohm.m layer above. When the well is approximately 3 ft
[0.9 m] TVD from the high-resistivity layer, all the phase-shift and attenuation measurements read
the local resistivity of 1 ohm.m. The depth of investigation of the propagation measurement
decreases as the local resistivity decreases and because the high-resistivity layer above does not
offer a path of lower resistance to the measurement currents, they remain in the 1-ohm.m layer.
1 Ohm-m
P16H
P22H
P28H
P34H
Phase
Resistivities
A16H
A22H
A28H
Attenuation
A34H
Resistivities
Figure 5-38. The proximity of a low-resistivity layer reduces the deeper attenuation
resistivities (bottom) before the shallower phase-shift resistivities respond (middle) as the
well (green line, top) traverses from high resistivity (brown) to a low-resistivity layer (blue).
In summary:
• Responses on the deep measurements not seen on the shallower measurements can be
due to proximity effects from a nearby layer.
Rattenuation
Depth (m)
Figure 5-40. The dielectric effect causes the shallow attenuation resistivities to read higher
than the deep attenuation resistivities. The phase-shift resistivities show little or no
separation.
In summary:
• An unusually high shallow attenuation response in a high-resistivity formation may be due
to the dielectric effect.
Attn Phase
Phase
Resistivity
s s
s Deep
s
Attn Phase
Deep Deep
Deep Attn
Deep Deep
s s s
Attn
Phase s
Phase
Attn Attn Deep
s
s Phase
s
Deep Deep Deep
Conductive Resistive Polarization Dielectric Conductive Resistive
Invasion Invasion & Anisotropy Shoulder Shoulder
Figure 5-41. Although single effects are summarized for resistivity separation, data may
respond to more than one effect. Phase-shift (blue) and attenuation (red) resistivities
range from shallow (vertex) to deep (base).
The selection of a resistivity tool for operations in the complementary region is based on the
following:
• sharp axial resolution required—laterolog measurement preferred
• imaging and dip determination required—laterolog measurement preferred
• greater depth of investigation required—propagation measurement preferred
• conductive invasion expected (Rxo < Rt)—laterolog measurement preferred
• resistive invasion expected (Rxo > Rt)—propagation measurement preferred.
2K
Laterolog Only
200
Rt
20
Complementary Region
• axial resolution
• images & dip Propagation Only
2 • depth of investigation
0.2
0.01 0.1 1.0 10.0 Oil
Based
Mud
Rmud filtrate / Rwater
Figure 5-42. Resistivity tool selection is charted for homogeneous, uninvaded formations.
Rmf = mud-filtrate resistivity.
Borehole
Invaded Formation
Uninvaded Formation
Rt
Rxo log
Rm Latero nse
o
Resp
Rm
Rx o
Rt
nse
espo
on R
agati
Prop
The reverse is true for resistive invasion. In this case propagation is the preferred measurement
because the induced currents preferentially measure the low Rt behind the resistive invaded
zone. The laterolog is severely affected by the high invaded zone resistivity in series with the
uninvaded zone and reads significantly higher than Rt.
The formation porosity can be determined provided that the fluid and matrix densities are known.
However, they are typically unknown, so additional measurements sensitive to the fluid (neutron
and magnetic resonance measurements) and lithology (neutron response, PEF, and
spectroscopy) are required. Knowledge of the formation lithology enables determining the grain
density.
Although formation density can be determined with a single GR source and detector, in most
logging tools two detectors are used to create a compensated density measurement, which
corrects for parallel standoff between the detectors and borehole wall. The detector closer to the
source is called the short-spacing detector; the one farther from the source is called the long-
spacing detector (Fig. 5-44).
standoff
formation
Figure 5-44. A dual-detector density measurement can compensate for the effect of limited
standoff between the detectors and the borehole wall.
The spine-and-ribs correction compares the density measurements of the short- and long-spacing
detectors (Fig. 5-45). If they read the same density they plot on the spine and no correction is
ρls
∆ρ
Actual formation ρb
ρls−ρss 2.0 g/cc
Rib
∆ρ < 0
e ρmud > ρB
p in 1.0 g/cc Infinite standoff
S reading ρmud
ρss
Figure 5-45. The spine-and-ribs technique for density measurements corrects for standoff
between the detectors and the borehole wall.
Because gamma rays are relatively easily stopped, the density measurement is shallow (2 to 3 in
[5 to 8 cm]). However, it can be well focused, which makes it suitable for azimuthal
measurements and imaging. Contact with the formation is enhanced by the presence of a
stabilizer with special low-density windows installed to exclude drilling mud and improve gamma
ray transport to and from the formation. Stabilizers can affect the drilling tendency of the BHA, so
the decision to run stabilized or slick (without the density stabilizer) is a decision that should be
made in consultation with the directional driller.
Neutrons can be classified into three main categories according to their energy level:
• fast neutrons—energy in excess of 1,000 eV
• epithermal neutrons—energy between 10 eV and 0.4 eV
Figure 5-46. Neutrons are classified as fast (high and intermediate energies), epithermal,
and thermal.
Because neutrons have no electrical charge, they pass through the electron cloud and interact
with the nucleus. There are four main types of neutron interactions:
• Elastic scattering—A neutron that strikes an atom scatters and loses some of its energy
to the target nucleus. The amount of energy transfer from the neutron to the nucleus
depends on the mass of the nucleus (the bigger the nucleus, the less the transfer).
• Inelastic scattering—When a fast neutron strikes an atom, a portion of its energy goes
into exciting the target nucleus. The target nucleus de-excites and emits gamma rays
with an energy that is characteristic of the atom. The neutron is deflected and continues
moving at reduced velocity.
• Radiative capture—A thermal neutron is absorbed by the target nucleus, producing a
compound nucleus (isotope). The nucleus de-excites immediately, emitting gamma rays
with an energy that is characteristics of the atom.
• Activation—As described for radiative capture, an atom that results in a radioactive
isotope or by nuclear reaction absorbs a neutron. After a delay that is governed by the
half-life of the isotope, the new atom experiences radioactive decay, emitting gamma
rays and other particles.
The flux of fast neutrons continuously emitted by the neutron source travels out in all directions
into the formation. As the neutrons progress, they go through three different phases:
The three processes of slowing down, thermal diffusion, and thermal neutron absorption occur
relatively fast, so that within a few microseconds dynamic equilibrium is reached for which the
total number of neutrons absorbed is equal to the number emitted by the source. Because
hydrogen is the main moderator, the size of the neutron cloud is determined by the formation
hydrogen index.
Formations with a high HI result in a small neutron cloud and hence the count rate at the
detectors is low. With the exception of gas, which has a low HI, formation liquids (water and oil)
contain similar amounts of hydrogen per unit volume. Thus HI can be related to porosity.
where
φ = true formation porosity.
Secondary effects refer to influences such as thermal neutron capture and differences in the
neutron response to various lithologies. These must be taken into account when deriving the
formation porosity from the neutron response.
The neutron response can be determined with a single neutron source and detector, but most
logging tools employ two detectors to create a compensated neutron measurement, which
reduces the effect of the borehole. Detectors closest to the source are called near detectors.
Detectors farther away from the source are called far detectors. The ratio of the near to far count
rates is used to determine the size of the neutron cloud and ultimately the formation porosity.
The transform from count rate ratio to neutron response differs from one lithology to another
because the scattering and capture cross sections of various lithologies differ. Because of this
variation, neutron “porosities” are quoted either as a limestone, sandstone, or dolomite neutron
porosity. The appropriate lithology should be selected for interpretation.
Even after borehole compensation, borehole effect is still the major environmental correction
required for the neutron response. The impact of the significant volume of hydrogen and chlorine
in the borehole fluid should be corrected before the measurement is used for interpretation.
Even after the neutron response has been corrected for environmental effects, the response is
primarily related to the HI of the fluid in the formation, so to derive formation porosity the
corrected neutron response must be divided by the fluid HI:
Zero p.u.
limestone
Oil filled
18 p.u limestone
Gas filled
18 p.u limestone
“Shale”
Figure 5-47. Separation between the density and neutron measurements presented on a
lithology-compatible scale (limestone in this example) reveals whether the formation is a
clean, freshwater-filled layer of the selected lithology.
With both the density and neutron measurements plotted on a limestone-compatible scale, the
two curves overlie when a clean, water-filled limestone of any porosity is encountered, as shown
in the top two layer rows of Fig. 5-47.
If the fresh water is replaced by oil, the density measurement decreases because the oil has a
lower density than the water it replaces. The neutron is also likely to decrease slightly because
the HI of the oil can be slightly less than water. This creates the “hydrocarbon separation” in the
third row of Fig. 5-47.
If the pores are filled with gas the separation becomes greater. The formation density is
significantly lower because the density of gas is much lower than that of water. The neutron
response is also significantly lower because the HI of gas is very low. Even though methane
(CH4) has more hydrogen per molecule than water or oil, the molecules in gas are much farther
The fifth row of Fig. 5-47 returns to the base case of freshwater-filled 18-pu limestone, where the
curves overlie.
If the fresh water is replaced by salty water, a reverse separation is created. The density of water
with dissolved salt is greater than that of fresh water, so the density measurement increases. The
presence of salt in the water has two effects on the neutron response. The dissolved salt pushes
the water molecules apart slightly, resulting in a small decrease in the HI. This would slightly
reduce the neutron response. However, salt contains chloride ions, which have a very large
capture cross section (Table 5-4). As previously discussed, the decreasing neutron count of the
thermal neutron measurement transforms to increasing neutron response. The presence in the
water of chlorides that capture neutrons means that fewer neutrons arrive at the detector to be
counted. The reduced count rate drives the neutron response higher, resulting in a separation
between the density and neutron responses, as in the sixth row of Fig. 5-47
The separation that occurs in shale depends on a number of factors, which vary among the
various types of shale. In general shale has a higher matrix density than limestone, resulting in
the density measurement reading higher. The neutron response in shale is high owing to the
presence of elements with a high capture cross section and effects on the neutrons because of
the high density of the shale. Because the neutron response in shale is complicated by a number
of factors, typical neutron responses are not supplied for shales in analytical charts or tables.
However, a large separation caused by both the neutron response and density reading high (Fig.
5-47), is generally attributable to the presence of shale.
The eighth row of Fig. 5-47 returns to the base case of freshwater-filled 18-pu limestone, where
the curves overlie.
If the formation is filled with fresh water but the matrix is sandstone rather than limestone, the
separation in the ninth row of Fig. 5-47occurs. The density decreases because the matrix density
3 3
of sandstone (2.65 g/cm ) is lower than the matrix density of limestone (2.71 g/cm ). Because the
porosity remains at 18 pu and filled with fresh water, the HI does not change from the limestone
case. The change in the neutron response is due to the differing scattering and capture cross
sections of the elements in the matrix. In the case of sandstone, more neutrons get through to the
detectors than in a limestone of the same porosity. Increasing neutron count results in decreased
neutron response.
If the matrix is dolomite rather than limestone, the effect is reversed. The dolomite has a higher
3
matrix density (2.85 g/cm ) than that of limestone so the measured bulk density is increased. As
with the sandstone case, the HI of the formation does not change but the dolomite matrix lets
fewer neutrons through than limestone does, so the neutron response is increased in dolomite.
3
To create a sandstone-compatible scale, the density scale should be changed from 1.65 g/cm to
3
2.65 g/cm and the sandstone lithology neutron response displayed rather than the limestone
lithology neutron response used in this example.
First, the tool's permanent magnetic field aligns, or polarizes, the spin axes of the protons in a
particular direction. This process, called polarization, increases exponentially in time with the
longitudinal relaxation time constant, designated as T1.
Next, the tool's oscillating field is applied to tip the protons away from their new equilibrium
position. The protons precess around the magnetic field in the same way that a child’s spinning
top precesses in the Earth’s gravitational field. Precession occurs as a body rotating about one
axis slowly rotates around a second axis. In the NMR case, this second axis is the static magnetic
field (Fig. 5-48).
Figure 5-48. During spin precession a proton (left) spins around the blue axis, which
precesses around the magnetic field (vertical). Similarly, the toy top (right) spins around
the blue axis, which precesses around the gravitational field (vertical).
The initial signal received from the aligned protons after the first oscillating field pulse is
proportional to the number of hydrogen nuclei associated with the fluids in the pores. This
amplitude is calibrated to provide the HI of the formation.
A sequence of pulses is transmitted by the tool to realign the protons while acquiring a series of
magnetic echoes from the spinning protons. The protons begin tipping back toward the original
direction in which the static magnetic field aligned them. In NMR terminology, this tipping-back
motion is called transverse relaxation, designated as T2. Relaxation results in the protons
becoming desynchronized with the remaining proton population. This causes a reduction in the
amplitude of the magnetic signal received by the tool’s antennas. The rate of decay of the
magnetic signal is the fundamental measurement of NMR logging tools and is controlled by the
three relaxation mechanisms:
• Surface relaxation—relaxation caused by the interaction of the protons with the walls of
the pores
• Bulk relaxation—relaxation caused by the interaction of protons with each other
• Diffusion relaxation—apparent relaxation caused by the aligned protons moving out of
the volume of investigation of the tool during the measurement sequence.
As summarized in Fig. 5-49 and Table 5-x, the magnetic resonance sequence involves an initial
wait time during which the protons in the formation fluids align to the tool’s permanent magnet.
This polarization alignment is exponential and is controlled by surface and bulk relaxation of the
protons. Diffusion relaxation does not play a part in polarization because the oscillating field has
not yet been used to create a resonant measurement volume.
0.8
t
− Decay rates provide
TT2
0.6 e information on pore and
fluid types.
0.4
0.2 Polarization, T1
Transverse Relaxation, T2
0
0 2 4 6 8 10 12
Time (seconds)
Figure 5-49. Magnetic resonance measurement involves the polarization of hydrogen
nuclei followed by a sequence of realignments to measure the transverse relaxation rates,
T2, of the nuclei.
Table 5-x. Control of Relaxation of the Polarization Rate, T1, and Decay Rate, T2
T1 Relaxation T2 Relaxation
Surface relaxation Yes Yes
Bulk relaxation Yes Yes
Diffusion relaxation No Yes
The T2 decay rate measured by the tool is the cumulative sum of multiple decay rates (Fig. 5-50,
top). The decay rate measured by the tool is inverted to find the component decay rates (Fig. 5-
50, bottom). This process generates a T2 distribution from fast decay rates on the left to slow
decay rates on the right. The total area under the T2 distribution equals the HI of the formation.
The height of the distribution at a given T2 number represents the proportion of the signal
decaying at that T2 rate.
Magnetic resonance analysis on cores is used to determine the T2 cutoff above which fluid flows
(Fig. 5-52). Fluid below the cutoff is called bound fluid, and it is not produced when the well flows
because it is either adsorbed to the surface of clay (clay-bound water) or trapped by capillary
pressure effects (capillary-bound fluid).
Figure 5-52. The T2 distribution before (blue) and after (red) centrifuging a core sample
shows a decrease in the T2 components to the right that results from the free fluid being
spun off and its signal lost. The nonmobile, or bound, fluid remains in the small pores.
With this type of core data, a T2 cutoff between free and bound fluids can be determined.
A T2 cutoff of 33 ms is generally used in sand-shale sequences. The T2 cutoff for carbonates is
not as consistent as in sand-shale lithology because the surface relaxation coefficient (related to
the paramagnetic content of the matrix) varies widely in carbonates.
Although NMR tools do not directly measure formation permeability, the pore-size information is
useful in transforms that correlate the T2 distribution to permeability measured on cores. To
quantify the position of the T2 distribution, either the bound fluid to free fluid ratio or the
logarithmic mean of the distribution (T2 log mean, T2LM) is used. For well placement purposes,
the key element is that the farther the T2 distribution is to the right, the larger the pores and the
higher the permeability.
An example of the use of magnetic resonance information for geostopping is shown in Fig. 5-
4
53Figure 5- (Rose et al., 2003 ). In this Middle East carbonate the total porosity does not change
significantly across the reservoir, as indicated by the relatively constant density (red), neutron
(blue), and total magnetic resonance porosity (black) responses in Track 2. However, the size of
the pores reduces on the flanks of the reservoir, which significantly reduces the permeability to
the point that the formation is no longer commercially productive. Drilling is stopped so that
valuable rig time is not wasted extending the well into nonproductive formation. The transition
4
Rose, D., Hansen, P.M., Damgaard, A.P., and Raven, M.J.: “A Novel Approach to Real Time
Detection of Facies Changes in Horizontal Carbonate Wells Using LWD NMR” Transactions of
th
the SPWLA 44 Annual Logging Symposium, Galveston, Texas, USA (June 22–25, 2003), paper
CCC.
Figure 5-53. NMR logging was used to guide geostopping based on the reduction in pore
size and resulting significant reduction in permeability. Courtesy of D. Rose.
(Can we find another log to replace 5-53 that is not copyright SPWLA? Another option is to get
the original data and reformat, showing slightly different interval and tracks.- This is not SPWLA
copyright as it is from the D. Rose presentation not from the paper)
Figure 5-54 shows a sequence of formations with their corresponding magnetic resonance
responses. The shale at the bottom of the formation column on the left has small pores and
hence strong surface relaxation. This condition results in fast polarization and echo decay of the
water in the shale. Consequently, the T2 distribution is to the left of the T2 cutoff, indicating that
the fluid is not free to move.
The water sand has both small and large pores that are water filled. The water in the large pores
may not interact with the pore walls and decay is at the bulk relaxation rate for water. Water bulk
relaxation is caused by hydrogen protons in the water interacting with other hydrogen protons in
other water molecules. The water in the small pores has significant surface relaxation and
polarizes and decays quickly. The spread of the T2 distribution for the water sand indicates the
range of relaxation rates. Some of the distribution lies above the T2 cutoff, indicating that some of
the fluid (in the center of the large pores) is free to move. Water in the small pores and around the
surface of the large pores is held in place by electrochemical and capillary effects and does not
flow, as indicated by the proportion of the T2 distribution below the T2 cutoff.
Gas Sand
Diffusion ~ f(TE,G,D)
T2 Cut Off
The gas sand is similar to the oil sand, with the oil replaced by gas. The water is displaced to the
edges of the pores and undergoes significant surface relaxation. The gas in the center of the
pores does not interact with the pore walls. Because gas molecules are generally small and
widely spaced, they do not interact frequently with each other, resulting in low bulk relaxation
effects. With low surface and bulk relaxation, the polarization of gas is considerably slower than
Figure 5-55. The acoustic arrivals of the sonic pulse that are of greatest interest are the
formation compressional, shear, and Stoneley arrivals.
An array of equally spaced receivers measures the acoustic wave train (Fig. 5-56). The distance
between the receivers is known, so measuring the time difference between an arrival at one
detector and the next allows the acoustic traveltime per unit distance (in units of seconds per
meter), otherwise known as the slowness, to be determined. The inverse of the slowness is the
acoustic velocity through the formation (in meters per second).
Receiver
Receivers
spacing
Transmitter
The wave train comprises various arrivals for which the velocity of each can be determined by
measuring arrival time at each of the detectors. The calculation process is called slowness-time
coherence (STC). The coherence of the waveforms is calculated (effectively, points of greatest
similarity between the waveforms are identified) and plotted as a color (red indicates highest
coherence, blue lowest) on a slowness versus time crossplot (lower left in Fig. 5-57). Time refers
to how long the arrival takes to propagate from the transmitter to receiver. Slowness refers to the
inverse of the velocity across the receiver array. The slower the arrival propagates through the
formation, the longer it takes to get to the receivers. This logic allows processing limits to be
applied (white dashed lines) to exclude computing physically unlikely velocities from any outlying
coherence peaks.
Receiver
spacing
Well Depth
Slowness
Time
Figure 5-57. STC processing extracts the velocity of the individual arrivals at each depth.
The data is then presented as a slowness versus depth log colors shading from blue (low
coherence) to red (high coherence).
STC processing enables computing the compressional, shear, and Stoneley velocities. Figure 5-
57 shows how the coherence (red to blue shadings) and the detected arrival peaks (black lines)
are plotted on a depth log.
The compressional velocity is used for time-to-depth correction of the seismic data and porosity
determination. The higher the porosity, the slower the compressional arrival through the
formation. Gas has a lower acoustic velocity than oil, which is in turn slower than water. The
acoustic velocity of the fluid in the pores must be taken into account when computing the porosity.
An interesting feature of porosity derived from acoustic logs is its insensitivity to isolated porosity.
If a pore is unconnected to its neighbors, it is surrounded by solid rock. The acoustic wave
propagates through solid material faster than through liquid, so the velocity measured at the
receivers is unaffected by unconnected pores. If the pore is connected, then the acoustic energy
is slowed because it has to cross the fluid in the throats connecting the pores. Acoustic porosity
shows a deficit to a nuclear porosity such as density-derived porosity (which measures all
porosity, whether it is connected or not). The difference between them is related to the isolated
porosity. Isolated porosity affects nuclear and resistivity logs but does not allow flow, creating
interpretation complexity if it is not quantified.
The compressional and shear velocities are used for calculating rock mechanical properties,
which have wide application in optimizing drilling, ensuring hole stability, and inputting to
mechanical earth models (MEMs) for reservoir behavior under changing pressure and stress
regimes.
Table 5-5. Photoelectric Factor and Volumetric Photoelectric Factor for Common
Lithologies
Dolomite 3.1 9
Anhydrite 5.1 15
Gamma rays are relatively easily stopped by dense material. This is particularly true of the low-
energy gamma rays involved in the photoelectric effect. The PEF measurement is the shallowest
of all LWD measurements and is highly sensitive to borehole and mud conditions. The
measurement is not borehole compensated like the density measurement is but is made with a
single detector (the short-spacing gamma ray detector in the density section). In light mud weight
with good borehole conditions, the PEF measurement can be used for lithology identification.
However, careful quality control is required in rugose boreholes or where heavy mud is used to
ensure that the PEF reading is not affected by the borehole environment.
The PEF is an important indictor of formation lithology. Knowing the lithology is required so that
the appropriate matrix density can be used in calculating the formation porosity from the density
measurement.
The production of gamma rays from neutron interactions involves one of two types of reactions:
• high-energy inelastic reactions
• thermal capture reactions.
For high-energy inelastic reactions, a high-energy (several MeV) neutron interacts with a nucleus
in the borehole or the formation. As a result, the nucleus emits one or more gamma rays that are
characteristic of the particular element (isotope). Inelastic gamma rays are important mainly for
carbon/oxygen (C/O) logging. For thermal capture reactions, a slow (thermal) neutron is absorbed
by a nucleus in the formation, borehole, or tool. While neutron absorption is unlikely at high
neutron energies, it occurs quite readily at thermal energies, especially in the presence of thermal
neutron absorbers such as chlorine, boron, and gadolinium. When neutron capture occurs, the
nucleus is often left in an excited state. When the nucleus de-excites from an excited state to its
ground state, high-energy gamma rays are emitted. These gamma rays, referred to as prompt
capture gamma rays because they are emitted immediately after neutron capture, are
characteristic of the isotope that captured the neutron. Many of the gamma rays emitted are very
energetic and scatter throughout the formation and wellbore, resulting in a cascade of associated
lower energy gamma rays.
When struck by an incoming gamma ray, the detector material "scintillates," which means that it
generates light, which is extracted from one end of the crystal. The scintillation light is absorbed
by the photocathode material in the photomultiplier tube. The photoelectric emission of an
electron is followed by electron multiplication to increase the magnitude of the signal. The result is
a composite gamma ray energy spectrum, which provides a measurement of the contribution of
various elements in the tool, borehole, and formation (Fig. 5-58).
The measured spectrum is the linear summation of gamma ray signals from elements in the
formation, borehole, and tool. The fraction of the spectral area resulting from gamma rays from a
particular element is called the relative elemental yield. In most well logging situations, the main
contributions to the measured capture gamma ray spectra come from the following elements: H,
Cl, Si, Ca, Fe, S, Ti, Gd, K, Mg, Cr, Ni, and Ba. Radiative capture by these elements results in the
emission of characteristic gamma rays unique for each element. The presence of an element is
inferred when the characteristic gamma rays are observed in the measured spectrum.
100
10
Silicon
Calcium
Iron
1
Tool Background
Chlorine
0.1
The quantities derived from spectral processing are called relative elemental yields. They can be
used to determine the ratio of one element to another in the formation (for example, Ca/Si) but do
not directly deliver the absolute concentration or weight fraction of any element. The next step is
to convert the relative elemental yields to elemental concentrations that can then be used for
quantitative interpretation. This is accomplished by using an oxides closure model, which takes
its fundamental justification from the knowledge that in sedimentary rocks, most elements exist in
their oxide forms and the sum of the rock-forming oxides is 1. In terms of the elements, this is
expressed as
+
SiO + TiO2 + Al2O3 + Fe2O3 + MgO + CaO + Na2O + K2O + CO2 + P2O5 + H2O + SO2 = 1. (5-20)
Although only a subset of these rock-forming elements is available in the form of relative yields
from the capture spectroscopy measurements, it is possible to adapt this equation to an oxides
closure model that ultimately enables conversion from yields to concentrations. Several factors
must be taken into consideration. First, unmeasured elements must be accounted for by relying
on natural correlations between elements in sedimentary rocks. For example, for environments
where aluminum can not be measured directly, an empirical relationship was developed to
compute Al from the elements silicon, calcium, and iron. The complementary nature between
Second, because the input measurements are relative yields, the measurement sensitivity of
each element must be accounted for. Some elements are more easily detected than others. Each
tool has sensitivity coefficients that quantify how easily each element is detected. For one
spectroscopy tool, calcium has a sensitivity coefficient of about 1.6 compared with 1 for silicon.
This means that a calcium atom can be detected 60% more easily in the spectra than a silicon
atom.
Third, not all of the yields measured come from the rock. For example, the analyzed yields
include tool background, as well as hydrogen and chlorine from both borehole and formation
fluids. To get just the rock elemental concentrations, the undesired yields are left out of the oxides
closure model, and the remaining yields are normalized to unity:
Y Y Y Y Y
F X Si Si + X Ca Ca + X S S + X Ti Ti + X FeAl Fe = 1, (5-22)
S Si S Ca SS S Ti S Fe
where
F = normalization factor that compensates for the elimination of Cl, H, and tool background from
the analysis and that the yields are relative and divided by their sensitivities
Yi = relative yield of element i
Si = sensitivity of element i to the prompt neutron capture measurement
Xi = oxide association factor used to convert the element to its appropriate oxide or oxide and
related elements. For Si and Ti, this term simply converts Si to SiO2 and Ti to TiO2. For calcium,
the model assumes that Ca primarily resides as CaCO3, thus accounting for both the CaO and
CO2 terms in Eq. 5-20.
Using the oxides closure model, it is possible to compute the weight fraction, W, of each element
in the formation from the relationship
Yi
Wi = F . (5-23)
Si
The resulting dry-weight elemental concentrations can be used to compute both mineralogy as
well as matrix properties including matrix density, matrix sigma, and matrix neutron responses.
For example, matrix density can be directly approximated as a linear combination of the elements
3
silicon, calcium, iron, and sulfur with a standard error of only 0.015 g/cm according to the
relationship
ρ matrix
= 2.62O + 0.0490 Si + 0.2274 Ca + 1.993 Fe + 1.193 S, (5-24)
where Si, Ca, Fe, and S are weight fractions of the elements silicon, calcium, iron, and sulfur as
derived from the oxide closure processing. In pure quartz, substituting a value of 0.47 for the
3
silicon weight fraction produces a matrix density of 2.65 g/cm .
The dry-weight elemental concentrations can also be used to calculate formation mineralogy
through the solution of simultaneous equations or converted into the major mineral groups based
on the SpectroLith* processing methodology (Fig. 5-59). Standard SpectroLith processing without
the use of Mg delivers
based on the measured silicon, calcium, iron, and sulfur dry-weight elemental concentrations
along with the derived aluminum concentration. The interpretation is based on an extensive, high-
quality core database comprising chemical concentrations and Fourier transform infrared
spectroscopy of the mineralogy measured on hundreds of sedimentary rocks.
The use of elemental capture spectroscopy and SpectroLith processing to determine the
formation mineralogy is especially useful because it is conducted without reference to any other
measurement and knowing the porosity is not required for the solution.
Elemental
Relative
Yields
Matrix Properties
e.g. grain density from the
elements
For well placement, individual elemental concentrations can be used for geochemical correlation
between wells and as layer markers. Quantitative lithology volumes from capture spectroscopy
significantly simplify porosity and subsequent saturation evaluation.
Figure 5-60. Sigma, the macroscopic thermal neutron capture cross section of the
formation, is closely related to the formation salinity.
The sigma measurement is acquired by measuring the count rate of either neutrons or gamma
rays in a single detector as a function of time. Initially the decrease in the count rate is due to the
effect of the tool and borehole in proximity to the detector. The borehole effect is relatively small
for sigma acquired using an LWD tool because the large LWD collar displaces most of the mud in
the borehole. Sigma, Σ, expressed in capture units (cu), is related to the decay rate of the neutron
or gamma ray counts in the detector:
4,550
Σ= ,
τ
(5-25)
where
τ = decay constant of the quasi-exponential decay of the neutrons or gamma rays (υs).
Typical sigma values for common formation components are in Fig. 5-61. There is a large
difference between the sigma values for hydrocarbons and salty water.
Sandstone = 4.3
Dolomite = 4.7
Calcite = 7.1
Lithology
CLAYS
Anhydrite = 12
∑ 0 5 10 15 20 25 30 35 40 45 50
Figure 5-61. Typical sigma values for common formation components show a large
difference between values for hydrocarbons and those for saline water.
which can be transposed into a simple linear equation for water saturation:
(Σ − Σ grain ) + φ (Σ grain − Σ HC )
Sw =
bulk
, (5-28)
φ (Σ water − Σ HC )
where
Sw = formation water saturation
φ = formation porosity
Σbulk = measured bulk formation capture cross section
Σwater = capture cross section of the water
Σgrain = formation solid (grain) capture cross section
ΣHC = hydrocarbon capture cross section.
In shaly formations sigma tends to correlate with the GR measurement as a result of the high
capture cross section of clays. In clean carbonates it tends to anticorrelate with resistivity (Fig. 5-
62, with the resistivity scale increasing to the right and the sigma scale increasing to the left). In a
formation filled with salty water, such as that shown in the lower part of Fig. 5-62, sigma reads
high because of the presence of chlorides in the water and the resistivity reads low owing to the
conductivity of the salty water.
10 ft
Res
oil
Sigma
water
Figure 5-62. Sigma and resistivity anticorrelate in clean formations where hydrocarbons
have displaced salty water. In the water interval, sigma is high and resistivity low. In the oil
interval, sigma is low and resistivity high. The upper interval (orange) shows a low-
resistivity pay zone in which the volumetric sigma measurement gives a better indication
of the formation water saturation because it is not influenced by the path-of-least-
resistance effect that complicates resistivity interpretation.
Sigma can be useful for well placement in low-resistivity pay zones (such as where resistivity
anisotropy complicates conventional saturation determination) because saturation can be
determined without use of resistivity measurements. Sigma can also be useful where shoulder
Figure 5-63. Acquiring a formation pressure measurement requires a seal between the tool
and formation, a piston to reduce the pressure in the tool below formation pressure so
that fluid flows into the tool, and a pressure gauge to measure the pressure response.
The sequence of formation pressure drawdown and buildup is called a pretest because it was
originally designed to confirm a pressure seal between the tool and formation prior to taking
further tests and samples. Today pretests are commonly used to acquire formation fluid pressure
and mobility data only (Fig. 5-64). The drawdown and buildup rates provide information about
how freely the formation fluid moves. The mobility of the fluid is defined as the ratio of the
formation permeability divided by the fluid viscosity. This takes into account the ease with which
the fluid moves (viscosity) and the ability of the rock to transmit fluids (permeability).
As shown in the example in Fig. 5-64, the pressure gauge initially reads the mud pressure in the
borehole. There is a slight increase as the mudcake on the borehole wall is compressed by the
sealing packer. The pressure in the flowline then drops rapidly as the pretest piston retracts. A
small pressure anomaly occurs as the pressure in the flowline drops below formation pressure
and the mudcake collapses into the flowline. At this point formation fluid begins to flow into the
flowline and continues to do so when the pretest piston stops. The pressure builds until the
Tool Sets
Pressure
Fluid Draw-down
Fluid Draw-down
(Measurement)
(Investigation)
Build-up 2
Build-up 1 Formation
Pressure
time
Time allocated for pressure measurement
Figure 5-64. A typical formation pretest comprises drawdown and buildup sequences that
are used to determine fluid mobility and pressure.
p = ρ fluid g n (TVD ),
(5-29)
where
ρfluid = fluid density
gn = gravitational acceleration
TVD = true vertical depth.
One of the major applications of pressure data is for identification of the fluids and their contacts
within the formation. In plots of pressure against the TVD of a well, fluids of similar density fall on
a line (Fig. 5-65). The gradient of the line is related to the density of the fluid, which can be used
to identify the type and density of the mobile fluid in the formation. Water, with a density about 1
3
g/cm , has the highest gradient, whereas oil and gas plot at lower gradients depending on their
composition, temperature, and pressure. The intersection of the gradients can be used to infer
the location of the fluid contacts in the formation, even if the well does not cross the contact.
Formation pressure data between wells is used to identify reservoir compartments and uneven
depletion.
Gas
Gas-Oil Contact
X600
TVD (ft)
Oil
Oil-Water Contact
X700
Water
Formation pressure can be used to drill horizontal wells with greater accuracy. Given the
uncertainties in surveying, wells that are intended to be horizontal may not actually be drilled
horizontally. By using measurements of the formation pressure, adjustments to the well trajectory
can be made to maintain a constant measured formation pressure, which indicates that he well is
being drilled at a consistent TVD and is therefore horizontal.
Source
MWD telemetry
Sensors
seismic reflector
Figure 5-66. Seismic while drilling can be configured as drillbit seismic service (left), which
uses vibrations from the drilling process as a seismic energy source. The drillbit seismic
service can be used only under specific conditions. The more widely used seismic LWD
system uses a conventional seismic source at surface and detects the signal downhole.
This deep directional capability bridges a gap in the real-time measurement portfolio (Fig. 5-67).
Seismic data is able to visualize large structural features, but because of the limited frequency
content of surface seismic data, some features are below seismic resolution. At the other end of
the scale, conventional LWD images indicate the presence of a formation boundary, but because
of their limited depth of investigation it is often too late to change trajectory in time to avoid exiting
the reservoir. Conventional propagation resistivities have been used successfully for well
placement relative to a boundary, but because these measurements are nonazimuthal, they do
not indicate the direction from which a boundary is approaching the wellbore and hence do not
10000
Surface seismic
1000
Seismic while drilling
100
Vertical Resolution (ft)
PeriScope UD/XD
10
1 PeriScope
Wellbore
0.1 Borehole images
0.01
Figure 5-67. PeriScope distance-to-boundary service delivers both directionality and depth
of investigation at a useful scale
The remote detection of surrounding layers opens up a whole new capability for well placement.
• Wells can now be positioned closer to the roof of a reservoir without risking exiting,
ensuring that the minimum attic oil is left between the wellbore and reservoir roof when
the reservoir is depleted.
• Wells can be drilled along meandering ancient river channels, detecting the banks, roof,
and floor of the sand to keep the wellbore positioned in the productive channel.
• Wells can be drilled parallel to fluid contacts, enabling the exploitation of thin oil rims, for
example.
• Events such as subseismic faulting can be identified and quantified. As thinner reservoirs
are drilled, the evaluation of subseismic faulting becomes increasingly important because
the borehole can exit the reservoir on encountering a fault.
• Evaluation of reservoir compartmentalization is improved through the ability to track the
top and bottom of layers, thereby improving estimation of the volume of hydrocarbons in
place.
• Determination of the resistivity in adjacent layers enables tracking water movement
remotely. Experience has shown that a single well can track waterflood fronts in multiple
layers simultaneously, reducing the number of observation laterals required to track water
movement.
One of the key capabilities of remote boundary detection is providing early warning of changes in
formation dip. Figure 5-68 shows a horizontal well encountering a change in formation dip. If a
conventional propagation resistivity measurement with a depth of investigation of approximately 4
ft [1.2 m] is positioned 40 ft [12 m] behind the bit, the change in formation dip of 5.7° is not
identified until when the bit begins to exit the reservoir. Under the same conditions, the 15-ft
40 ft.
20.6 degrees
15 ft.
40 ft.
15ft
4 ft
PeriS Propa
cope gation
15 Resis
ti vity
Figure 5-68. Remote detection of a change in formation dip can provide sufficient
forewarning to enable trajectory adjustment and remain in the reservoir.
The increased transmitter-receiver spacings (34, 84, and 96 in [0.86, 2.13, and 2.44 m]) and
operation at lower frequencies (100 kHz, 400 kHz, and 2 MHz) than conventional propagation
resistivity measurements gains greater depth of investigation for the PeriScope measurement.
Both directional phase-shift and attenuation measurements are made.
The tilted receiver coils create directional sensitivity at a 45° angle into the formation. The
conventional axial coils do not show any sensitivity to tool orientation as the tool rotates. With
tilted coils the receiver voltage depends on the orientation of the coil with respect to conductivity
changes in the formation.
R3 T5 T3 T1 R1 R2 T6 T2 T4 R4
34”
84”
96”
Figure 5-69. PeriScope deep directional measurements are acquired at three frequencies
using axial propagation resistivity transmitters and directional receivers tilted relative to
the axis of the tool.
The sensitive volume of a paired axial transmitter and tilted receiver is shown in Fig. 5-70. One
side shows positive sensitivity and the other shows negative sensitivity.
Sensitivity
-0.0001
-0.001
-0.01
-0.1
0.1
0.01
0.001
0.0001
Figure 5-70. The sensitive volume of an axial transmitter and tilted receiver pair produces
positive (upper) and negative (lower) lobes.
When rotated in a homogenous layer, the responses from the positive and negative lobes cancel
each other, resulting in a net zero phase-shift and attenuation response. In a layered formation
the directionality created by the tilted receiver results in a sinusoidal response. Figure 5-71
compares the phase shift at a tilted receiver to that at an axial receiver caused by excitation from
an axial transmitter when the tool is placed close to a formation boundary. As the tool rotates, the
axial receiver voltage is constant but the tilted receiver voltage shows sinusoidal variation with the
tool rotational azimuth angle. The magnitude of the variation is related to the distance to the
conductive boundary, and the signal maximum indicates the direction of a more conductive
formation. This is the basic concept of the azimuthal measurements, providing both the direction
and distance to the formation boundary.
T R
Tool
Rotation
Tool Rotation
T
R
Figure 5-71. When placed near a change in formation resistivity, the signal in an axial
receiver from an axial transmitter does not change with rotation of the tool. The signal in a
tilted receiver varies sinusoidally with tool rotation. The magnitude of the sinusoid is
related to the distance to the resistivity change and the resistivity contrast between the
layers.
An additional benefit of symmetrization is that the sensitivity map becomes symmetric between
the transmitter and receiver, with the dominant contribution from the region between them (Fig. 5-
73).
=
T R
Unsymmetrized Symmetrized
directional attenuation (dB) directional attenuation (dB)
Rt = 3 ohm.m.
Rh=6 ohm.m.
TVD (ft)
TVD (ft)
Rv=30 ohm.m.
Rt = 1 ohm.m.
Sensitivity
-0.0001
-0.001
-0.01
-0.1
0.1
0.01
0.001
0.0001
Figure 5-73. Symmetrization results in symmetric sensitivity between the transmitter and
receiver.
60° 1 Ωm
10 Ωm
Amplitude
60°
or Phase
Figure 5-74. The symmetrized directional response is a maximum when the positive lobe is
oriented toward the conductive bed and a minimum when oriented away. The azimuth of
maximum amplitude found by sinusoid fitting enables determination of the azimuth of the
bed.
The amplitude of the response depends on the distance to the resistivity change and the
resistivity contrast:
• The closer the layer, the greater the amplitude.
• The higher the contrast, the greater the amplitude.
If the layers have the same resistivity or the bed with contrasting resistivity is beyond the depth of
investigation of the directional measurements, the amplitude of the sinusoid is zero.
The PeriScope directional response in the majority of cases is summarized in Table 5-6. The sign
of the response (positive or negative) is controlled by the product of the location of the change
above or below the tool and whether the bed is more or less resistive than the layer in which the
tool is located.
• Beds more conductive than the layer in which the tool is located create a positive response.
• Beds more resistive than the local layer create a negative response.
• Resistivity changes above the tool create a positive response.
• Resistivity changes below the tool create a negative response.
Conductive Resistive
+ –
Above
+ + –
Below
– – +
As shown in Table 5-6, a positive directional response can be due to a more conductive bed
above the tool or a more resistive bed below the tool. Similarly, a negative response can be due
to a more conductive bed below the tool or a more resistive bed above the tool. This is resolved
by using multiple directional measurements with differing depths of investigation.
The signal from a single bed can be thought of as the product of the conductivity contrast
multiplied by the proximity of the bed to the tool. The total directional signal amplitude measured
by the tool is the sum of the response from the positive and negative lobes. In situations where
the signals from above and below the tool have the same amplitude, the total directional signal is
zero because the positive and negative responses cancel.
These concepts can be explored by reviewing the tool response to the three-layer formation in
Fig. 5-75. In the bottom panel, the three layers of different resistivity are traversed by a well
(green line) inclined at 85° (5° incidence angle to the horizontal layering). The middle panel
shows the conventional propagation phase (green) and attenuation (blue) resistivity responses.
As each boundary is crossed, the propagation resistivities show positive polarization horns.
These horns are useful for identifying the location of the boundaries, but they do not indicate
whether the boundary is approaching from above or below.
The upper panel shows the PeriScope directional attenuation response. This response is not a
resistivity but an attenuation measurement in decibels. The maximum amplitude of the response
occurs where a boundary is crossed. Starting from the left, the attenuation signals are zero when
the tool is far from the boundaries. When the 20-ohm.m layer is within the depth of investigation
of the deeper attenuation measurement, the response becomes positive because the detected
layer is below the tool (negative) and more resistive (negative) than the local layer. Negative
multiplied by negative gives a positive response.
PeriScope
0 Attenuation
Measurement
-20 dB
2000 Ohm-m
Conventional
Propagation
Resistivity
0.2 Ohm-m
Measurement
2 Ωm
True Vertical Depth (ft.)
Formation
20 Ωm Resistivities
1 Ωm
Figure 5-75. The response to a three-layer formation model (bottom panel) of conventional
propagation resistivity (middle panel) shows positive polarization horns whereas the
PeriScope directional attenuation measurement (top panel) develops a maximum response
where a boundary is crossed, with the sign of the response controlled by the location of
the layer above or below the tool and whether it is more or less resistive than the layer in
which the tool is located.
The response becomes increasingly positive until the tool passes through the boundary. Below
the upper boundary, the response becomes less positive as the signals from above and below
the tool partially cancel each other. When the tool is in the 20-ohm.m layer, the 2-ohm.m layer is
relatively conductive (positive) and above the tool (positive) so the signal from the upper layer is
positive. The lower layer is also relatively conductive (positive) but because it is below the tool
(negative) the product of the positive and negative result in a negative signal from the lower layer.
When in the 20-ohm.m layer, the relative proportion of the total signal resulting from the positive
signal from above and the negative signal from below depends on how close the tool is to the
respective layers. Just below the upper layer, the net response is positive because the positive
signal from the upper layer dominates. Closer to the lower layer. the signal becomes negative as
the lower layer dominates. At some point in the layer, the positive and negative signals balance,
resulting in a zero net response.
The position of the zero response is slightly closer to the less conductive (the more resistive) of
the two layers because the signal from this layer is weaker than the signal from the more
conductive layer. For the two signals to be of equal amplitude and hence cancel, the tool must be
closer to the less conductive layer. In a layer between two layers of equal conductivity, the zero
point is in the geometric middle of the local layer.
As the tool passes into the 1-ohm.m layer, the signal amplitude remains negative because the 20-
ohm.m layer is more resistive (negative) and above the tool (positive), resulting in a negative
• Measurement naming
The conventional propagation resistivities from the axial transmitters and receivers are named in
the same manner as normal arcVISION resistivities. For example, the attenuation resistivity
acquired with 34-in transmitter-receiver spacing at 400 kHz is called A34L, and the phase-shift
resistivity acquired with the same spacing but at 2 MHz is called P34H.
The names of the directional measurements used for distance to boundary determination have
four components:
• symmetrized (S) or antisymmetrized (A)
• phase (P) or attenuation (A)
• shallow (S), medium (M), or deep (D)
• 100 kHz (1), 400 kHz (4), or 2 MHz (2).
The antisymmetrized data is used for anisotropy calculations and should not be used for distance
to boundary inversions until after further development.
The boundary orientation with respect to the top of the hole is called
• DANG—boundary orientation (azimuth) derived from the 100-kHz and 400-kHz
measurements
• HANG—boundary orientation (azimuth) derived from the 400-kHz and 2-MHz
measurements.
Resistivity anisotropy measurements derived from the transverse transmitter (T6) start with the
letters AN, followed by phase (P) or attenuation (A) and the frequency number (1, 4, or 2) as
previously listed. These anisotropy measurements should not be used in distance to boundary
inversions until after further development.
Because of symmetrization of the directional measurements, the wellbore incidence angle to the
layers and resistivity anisotropy of the beds above and below do not influence the measurements
and hence do not need to be known or determined as part of the interpretation.
The local layer resistivities, Rh and Rv, are determined by using conventional propagation
resistivity measurements available from the axial transmitters and receivers on the tool. Solving
for local-layer resistivity anisotropy requires both conventional phase and attenuation resistivity
measurements. The resistivity anisotropy in the local layer is only determined to account for its
It is recommended that one conventional attenuation and one conventional phase resistivity be
used in inversion processing. Multiple conventional phase resistivity inputs are not recommended
as the inversion model can only solve for anisotropy effect on the conventional measurements, it
cannot account for invasion effects. If the conventional phase resistivities separate due to
invasion effect the inversion model will try to explain the separation by increasing the apparent
formation anisotropy, resulting in an inaccurate assessment of the local layer resistivity.
Inaccurate local layer resistivity reduces the accuracy of the distance to boundary evaluation.
The remaining four unknowns must be solved for by using directional measurements. A minimum
of four directional measurements is required, though it is recommended that six be used if the
real-time telemetry bandwidth is sufficient.
Ru
Rh , Rv
hu
hd
Rd
Figure 5-76. PeriScope measurement interpretation is based on a three-layer model with
up to six unknowns.
5.14.2.4 Crossplots
Crossplots are a way of visualizing the sensitivity of the PeriScope measurements to various
conditions. They can be used to determine which of the array of measurements gives the best
sensitivity to the expected formation resistivities. Appropriate crossplots should be generated
before a PeriScope job so that they can be used in place of the automatic inversion if problems
arise with Internet connectivity or the computer running the inversion processing.
A crossplot is a 2D representation of the data, with four of the model unknowns fixed, which
leaves two to be presented on the crossplot. The most common crossplot configuration is to
assume that the resistivities of the beds above and below the well are known and that the local
layer thickness is also known. This reduces the problem to solving for the local isotropic resistivity
and the distance to one of the boundaries. The distance to the other boundary is known because
it is the bed thickness minus the distance to the opposite boundary.
Ru = 2 Ω-m.
Rh =, RRvv = ?
hu hu = ?
20 ft.
hd hd = 20 - hu
Rd = 1 Ω-m.
Figure 5-77. A formation model representing a 20-ft-thick isotropic reservoir layer between
an upper bed of 2-ohm.m resistivity and a lower bed of 1-ohm.m resistivity.
To keep the crossplot as simple as possible, a deep directional response is plotted against a
conventional propagation resistivity. The conventional measurement helps define the resistivity of
the local layer. The directional measurement is used to determine the distance to the boundaries.
Figure 5-78 shows one of many possible crossplots that can be made for the formation model in
Fig. 5-77 by using two of the multiple measurements from the PeriScope tool. The conventional
P28H resistivity is plotted on the x-axis because it is relatively shallow and hence responds
primarily to the resistivity of the local layer. The SAD4 directional attenuation measurement is
plotted on the y-axis because it is a deep measurement with good sensitivity to the adjacent
beds.
The solid lines represent lines of Rt. The dashed lines represent the distance to the upper
boundary. For example, if the measured P28H is 9 ohm.m and the SAD4 is 0.5 dB, plotting these
values on the crossplot (red dot) reveals that Rt equals 9 ohm.m (purple solid line) and the upper
boundary is 8 ft [2.4 m] (aqua dashed line) above the borehole. Alternatively the plot can be used
to deduce tool responses. For example, when 2 ft [0.6 m] from the upper boundary in a 29-ohm.m
reservoir (blue dot), P28H is 21 ohm.m and SAD4 is 7 dB. These measurements indicate that 2 ft
from the upper conductive layer, the phase 28-in [71-cm] transmitter-receiver resistivity measured
at 2 MHz is lower than the local resistivity because part of the response is coming from the lower
resistivity bed above. Getting any closer to the upper bed results in the P28H reading
considerably higher because it polarizes close to the boundary.
The distance to the lower boundary is the 20-ft thickness of the reservoir layer minus the distance
to the upper boundary. The 10-ft [3-m] distance to upper boundary corresponds to a negative
SAD4 response. This is because the lower bed is more conductive, which creates a stronger
response in the directional receivers. To have the signals from the upper and lower beds balance,
the tool must be slightly closer to the less conductive upper bed. In this case the directional zero
point occurs at a distance of approximately 9.5 ft [2.9 m] to the upper bed.
The crossplot shows that the distance to the boundary is only weakly related to the local reservoir
resistivity because the local resistivity has a range of only 5 to 50 ohm.m. If the local resistivity
equaled the upper layer resistivity of 2 ohm.m, the distance would become a strong function of
the local resistivity and ultimately the lines would collapse to a point because PeriScope service is
unable to determine the distance to the boundary if there is no resistivity contrast with the target
layer.
5 ohm-m
9 ohm-m
17 ohm-m
29 ohm-m
50 ohm-m
P28H (ohm-m.)
Figure 5-78. This PeriScope response crossplot is for the formation in Fig. 5-77.
Crossplotting responses is a valuable technique for helping understand how the tool responds in
a wide variety of conditions.
If data communication was lost to a rig running PeriScope service in the formation modeled in
Fig. 5-78, drilling could continue with the Schlumberger field engineer adjusting the trajectory
based on the SAD4 response. Depending on the desired distance to the upper boundary, the well
could be steered up if the SAD4 response dropped below a target value and steered down if the
SAD4 value became too large. For example, if the desired distance to the upper boundary was
9.5 ft, then the well could be steered to maintain an SAD4 value of zero. If the SAD4 response
became positive the well placement recommendation to the client would be to drop inclination
slightly. If the SAD4 became negative (indicating getting closer to the lower boundary) the well
placement recommendation to the client would be to build trajectory inclination slightly.
Ru
Rh, Rv
hu
hd
Rd
Figure 5-79. PeriScope inversion solves for the resistivity and distance to layers above
and below the wellbore based on a three-layer model (left). The resulting information is
displayed in real time as a color-coded resistivity cross section of the formation (right).
The inversion is calculated on a point-by-point basis. The results from one point are not used to
guide the results for subsequent points. The independence of the points means that trends are
most likely real and not a consequence of one point affecting the next.
As discussed previously, the PeriScope tool provides an azimuth to the boundaries around it,
based on the assumption that the layers above and below are parallel. The boundary azimuth, in
conjunction with the distance to boundary information, is presented in an azimuthal view (Fig. 5-
80Figure 5-). Generally the boundary positions derived from the last 10 inversion points are
displayed with the most recent data displayed in the brightest color. This allows evaluation of the
DTB trend along the trajectory.
The distance and azimuth information enables steering a well in both TVD and azimuth relative to
a resistivity boundary without having to come into contact with it. For example, in the situation
shown in Fig. 5-79, the well could either be turned up to avoid the lower boundary or turned to the
left, or a combination of the two depending on the most appropriate position for the wellbore in
the target layer.
20 dB
PeriScope
0 dB Attenuation
Measurement
-20 dB
2000 Ohm-m
Conventional
Propagation
Resistivity
0.2 Ohm-m
Measurement
2 Ωm
True Vertical Depth (ft.)
Formation
20 Ωm Resistivities
1 Ωm
Real-Time
Distance to
True Vertical Depth (ft.)
Boundary and
Resistivity
Inversion
Results
Figure 5-81. The automatic PeriScope inversion (bottom panel) solves for the three-layer
model that best explains the measured PeriScope responses.
Starting from the left, the inversion finds the lower resistive boundary, but the response is not
consistent because the directional signal from the tool is weak. The weak signal results from the
conductive formation surrounding the tool, which limits the depth of investigation of the induced
measurement currents in the formation. As the tool approaches the boundary, the inversion
shows the boundary clearly because the signal becomes stronger. While the tool is in the 2-
ohm.m layer the inversion cannot solve for the 1-ohm.m layer because the inversion is limited to
solving for one boundary on either side of the tool. For this reason the 20-ohm.m layer appears to
be infinitely thick when the tool is in the 2-ohm.m layer. Once the tool crosses into the 20-ohm.m
reservoir layer, the upper and lower boundaries of the reservoir layer are visible. The lower
boundary is not well defined at first because the signal from the 2-ohm.m layer dominates the
PeriScope response. As the tool moves toward the center of the 20-ohm.m reservoir layer, the
As the tool approaches the lower bed, the accuracy of the upper bed description deteriorates.
The conductivity of the lower bed dominates the PeriScope response when the tool is close to it.
Once the tool crosses the lower boundary of the reservoir, the upper boundary disappears
because the inversion cannot solve for more than one boundary on each side of the tool.
Although the location of the boundary between the lower bed and reservoir is reasonably well
defined, the resistivity value for the reservoir is not consistent once the tool is in the lower bed.
Again, the conductivity of the lower bed constrains the measurement currents around the tool.
The higher resistivity of the reservoir layer does not create an attractive path for the measurement
currents, which are seeking the path of least resistance. Thus the PeriScope measurements have
relatively little sensitivity to the resistivity of the reservoir layer and consequently the inversion
struggles to deliver a consistent value.
As can be seen from this example, the inversion response is closely linked to the physics of the
PeriScope measurements. Where the measurements do not have the sensitivity to define a
parameter, the inversion is not able to clearly define the reservoir. Understanding where the
PeriScope measurements and inversion can and cannot provide answers is a vital part of proper
prejob preparation.
• Challenging environments
The basis of the PeriScope inversion on a three-layer model results in several challenging
environments in which the inversion may deliver a cross section that is not a good representation
of the subsurface.
One challenging environment for PeriScope inversion is where more than three layers affect the
response of the tool. In thin layers, for example, the PeriScope measurements mix the responses
from the various layers, with a bias toward the more conductive layers because the measurement
currents seek the path of least resistance.
40 Ω 6Ω 6Ω
30 Ω
30 Ω Input
3Ω
100 Ω 100 Ω Model
70 Ω 7Ω
70 Ω
6Ω 8Ω
30 Ω
Inversion
3Ω 90 Ω
100 Ω Output
70 Ω 7Ω 70 Ω
Figure 5-82. In thin layers, the PeriScope measurements respond to more than three
layers, resulting in potentially misleading inversion results.
In the thin-layered formation model in Fig. 5-82, a horizontal well is drilled through a 100-ohm.m
layer, over which are a 3-ft- [0.9-m-] thick, 30-ohm.m layer and a 40-ohm.m layer. Below the 100-
ohm.m layer is a 70-ohm.m layer(left side of the top panel). When water-flooded, the four layers
In the left and middle portions of the bottom panel, the PeriScope inversion does a good job of
finding the correct boundary location and determining the resistivity on the other side of the
boundary. However, the 40-ohm.m layer (left) is not identified because the more conductive 30-
ohm.m layer dominates the signal. When the topmost layer is flooded, its resistivity drops to 6
ohm.m (middle and right). The PeriScope directional measurements respond to this conductivity
and average it in with the 3-ft-thick, 30–ohm.m layer. The location of the upper boundary is
biased toward the conductive layer (blue dots upper right), and the resistivities of the layers are
also biased. The inversion identifies the local layer as having a resistivity of 90 ohm.m rather than
100 ohm.m and the top layer as having a resistivity of 8 ohm.m rather than 6 ohm.m. In thin
layers care should be taken to understand how the PeriScope tool and inversion combination
responds to the layers and what boundaries can be identified.
Increasing Resistivity
Ramp
Resistivity
Profile
PeriScope
Tool
Measurement Volume
Figure 5-83. In a resistivity ramp profile, the PeriScope inversion represents the smooth
resistivity change with a three-layer approximation because this is the model it solves for.
The inversion can solve for a three-layer model only, so it represents the smooth resistivity
change as a three-layer sequence with higher resistivity above and lower resistivity below. If the
borehole trajectory moves up or down in the ramp profile, the boundaries above and below the
tool follow the trajectory. Inversion boundaries that appear to follow the well trajectory are a good
The number of resistivity boundaries in the formation is not specified in the probabilistic approach.
The multilayer inversion determines the probability of a boundary at each point above and below
the well within the depth of investigation of the PeriScope measurements. The probability
distribution is then plotted at subsequent points along the wellbore, creating a series of probability
“wiggle” traces, as shown by the vertical black lines in Fig. 5-84. The probability wiggle traces can
be interpreted in a manner similar to seismic wiggle traces. The location of a boundary is
established where there is good coherence between high probability peaks on the wiggle traces.
Color is used to identify the resistivity of the various layers, and the opacity of the color indicates
the confidence in the resistivity evaluation. The closer to white the color becomes, the lower the
confidence in the resistivity value at that point.
Figure 5-84. Probabilistic inversion techniques enable the detection of multiple boundaries
above and below the wellbore (thick black line).
Starting from the left of Fig. 5-84Figure 5- and following the well trajectory (thick black line) to the
right, five layers are identified. When the tool is located in the uppermost green layer, the lower
boundary of the layer is clearly defined while the positions of the three boundaries below the well
become increasingly uncertain (the probability peaks are lower amplitude and wider) with
distance from the well. The opacity of the color increases closer to the well, indicating increasing
confidence in the inversion-derived layer resistivities near the well. As the well traverses through
the layers, multiple coherent boundaries remain visible, with more sharply defined probability
distributions closer to the well. The probability distribution of the upper boundaries widens and
The application of the multiple-boundary inversion to the very deep directional measurements of
the next generation of distance to boundary technology yields unprecedented real-time
information about the geometry and resistivities of formation layering. Figure 5-85 shows an
example application. During landing, the top of the target reservoir zone (upper yellow layer) is
detected sufficiently early to allow adjustment of the trajectory to ensure that the well lands just
below the top of the target layer. This guidance eliminates the need for a pilot well to identify the
location of formation tops and ensures that the well landing in the zone is smooth, facilitating the
installation of casing and completions. By knowing the location of the well with respect to the
layering at all times, wellbore undulations are avoided. In turn, production is improved because
water accumulation in troughs in an undulating trajectory and subsequent choking of hydrocarbon
flow are avoided.
In addition to enabling steering the well at the top of the target layer, which minimizes attic oil
above the well, the deep directional measurements can be used to identify the location of the
oil/water contact (OWC). This facilitates well placement sufficiently far from the water and
provides valuable information for reservoir volumetric calculations. On crossing a fault, the ability
to identify multiple boundaries provides greater confidence in determining the location of the well.
Consequently, well placement decisions to adjust the trajectory to place the well back at the top
of the target layer can also be made with greater confidence. Information about the throw of the
fault and the OWC in the new fault block helps further refine reservoir volumetric estimates. In
Clearly a great amount of information can be extracted when multiple layers are illuminated
above and below a well. The information is useful not only for real-time well placement
optimization but also for static reservoir evaluation and dynamic reservoir production modeling.
• Look-ahead capability
Look-around, look-ahead (LALA) capability has long been on the wish list of well placement
professionals. The remote detection technologies outlined previously deliver look-around
capability but do not provide information about formation changes ahead of the bit. Ongoing
electromagnetic research suggests that look-ahead capabilities can be developed, enabling the
identification of layering and fluid changes before the bit enters the different formation. For
example, in low-inclination wells drilling can be stopped before an OWC is intersected. In high-
angle and horizontal wells, look-ahead capability would enable the identification of faults, allowing
TVD trajectory changes to be made in anticipation of fault throw or azimuthal trajectory changes
to avoid the fault if it is being approached obliquely.
Accurate well placement helps improve well construction efficiency, reduce drilling risk, extend
reservoir contact, maximize reservoir exposure, improve well performance, enhance ultimate
hydrocarbon recovery, and improve reservoir development economics. Are you on target?