You are on page 1of 113

Halliburton Overpressure

CONTENTS

INTRODUCTION
I.1 PRESSURE GRADIENTS
I.1.1 Normal Pressures
I.1.2 Conversion Constants
I.1.3 Equivalent Mud Weight (EMW)
Example

SECTION 1 PHENOMENA OF OVERPRESSURE


1.1 OVERPRESSURES - SUGGESTED MECHANISMS
1.1.1 Hydrocarbon Reservoirs
1.1.2 Undercompaction (Sedimentary Loading)
1.1.3 Aquathermal Pressuring
1.1.4 Aquifers
1.1.5 Charged Upper Sands
1.1.6 Tectonic Loading
Uplift
Faulting
1.1.7 Salt Diapirism
1.1.8 Mud Volcanoes
1.1.9 Clay Diagenesis
1.2 SUBNORMAL FORMATION PRESSURES
1.2.1 Artificial Production
1.2.2 Precipitation
1.2.3 Potentiometric Surface
1.2.4 Osmosis
1.2.5 Temperature Change
1.2.6 Epeirogenic Movements
1.2.7 Formation Foreshortening
1.2.8 Decompressional Expansion
1.2.9 Different Densities of Oil
1.3 PROBLEMS ARISING FROM OVERPRESSURE
1.3.1 Lost Circulation
1.3.2 Differential Sticking
1.3.3 Formation Damage
1.3.4 Maintaining Rate of Penetration
1.4 SUMMARY OF OVERPRESSURE IN THE NORTH SEA

SECTION 2 OVERBURDEN GRADIENTS


2.1 TERZAGHI AND PECK EQUATION
2.2 OVERBURDEN PRESSURE

i
Halliburton Overpressure

2.2.1 Overburden Stress


2.2.2 Water Depth and Air Gap
2.3 DETERMINATION OF OVERBURDEN STRESS GRADIENT
2.4 AGIP FORMULA FOR CONVERSION OF SONIC TO BULK
DENSITY

SECTION 3 dC EXPONENT
3.1 DEVELOPMENT OF THE DRILLING RATE EQUATION
3.1.1 Rotary Speed
3.1.2 Weight on Bit
3.1.3 Tooth Efficiency
3.1.4 Drilling Hydraulics
3.1.5 Differential Pressure (P)
3.1.6 Overburden Pressure
3.1.7 Pore Pressure
3.1.8 Drillstring Effects
3.1.9 Matrix Strength
3.1.10 Lithology
3.2 DRILLING RATE EQUATIONS
3.2.1 Bingham
3.2.2 Jorden and Shirley (1966)
3.2.3 Rehm and McClendon (1971)
3.3 QUANTITATIVE PRESSURE EVALUATION FROM THE dc
EXPONENT
3.3.1 Ratio Method
3.3.2 Equivalent Depth Method (Matrix Stress Equation)
3.3.3 Eaton Equation (1975)
3.3.4 Determination of b Exponent
3.3.5 Back-calculation of dcn
3.3.6 Construction of Eaton Overlay
3.4 INTERPRETATION OF THE dc EXPONENT
3.4.1 Shales/Claystones
3.4.2 Argillaceous Siltstones
3.4.3 Calcareous Claystones
3.4.4 Shifting NCT
Procedure for Shifting NCT
Bit Shifts
Ice Sheet Compaction
Unconformities
3.5 NOMOGRAPH FOR QUICK dC EXPONENT CALCULATION

SECTION 4 OTHER INDICATORS OF ABNORMAL PRESSURE


4.1 GAS LEVELS

ii
Halliburton Overpressure

4.1.1 Connection Gas


4.1.2 Trip gas
4.1.3 Other Factors Affecting Gas Readings
4.2 INCREASED DRAG AND TORQUE
4.3 MUD PUMP PRESSURE
4.4 PIT LEVELS
4.5 FLOW METERS
4.6 CUTTINGS
4.6.1 Abnormal Trip Fill Up
4.6.2 Shale Density
Methods for Wellsite Evaluation of Shale Density
Bulk Density
Density Column Method
4.6.3 Percent Montmorillonite (Shale Factor)
Shale Factor Determination
4.7 CHLORIDES
4.8 FLOWLINE TEMPERATURE
4.8.1 Geothermal Gradients
4.8.2 End-to-End Plots
4.8.3 Horner Plots and Bottom Hole Temperature (BHT)

SECTION 5 PRESSURE ESTIMATION FROM WIRELINE LOGS


5.1 EVALUATION OF PORE PRESSURE BEFORE DRILLING
5.1.1 Mud History and Drilling Reports
5.1.2 Geological Correlation
5.1.3 Geophysical Aspects
5.1.4 Wireline Logs
5.2 GAMMA RAY LOG
5.3 SONIC LOG
5.3.1 Principle of the Sonic Tool
5.3.2 Cycle Skipping
5.4 RESISTIVITY
5.5 DENSITY
5.5.1 Formation Density Log (FDC)
5.5.2 Density/Neutron (FDC/CNL)
5.5.3 For All Logs
5.5.4 Density/Sonic Crossplot
5.5.5 Using Data Quantitatively
For Sonic Data
5.6 FORMATION PRESSURE TESTS (RFT/FMT)

SECTION 6 OVERPRESSURE DETECTION - SEQUENCE FOR WELLSITE


ANALYSIS

iii
Halliburton Overpressure

6.1 BEFORE DRILLING


6.2 PRESSURE EVALUATION DURING DRILLING
6.3 AFTER DRILLING

SECTION 7 FRACTURE PRESSURE


7.1 FRACTURE GRADIENTS
7.1.1 Mechanism of Formation Fracturing
7.2 LEAK OFF TESTS AND FRACTURE PRESSURES
7.2.1 Hubbert and Willis (1957)
7.2.2 Poisson's Ratio
7.2.3 Matthews and Kelly Method
7.2.4 Eaton (1969)
7.2.5 Anderson et al (1973)
7.2.6 Daines (1982)
7.2.7 Poisson's Ratio Calculations

iv
Halliburton Overpressure

SECTION 1
PHENOMENA OF OVERPRESSURE

1.1 OVERPRESSURES - SUGGESTED MECHANISMS


For overpressures to develop and be maintained, fluid flow must be inhibited or prevented.
For this to occur both vertical and lateral seals are required.

1.1.1 Hydrocarbon Reservoirs


In sealed reservoir rocks (eg lenticular reservoirs, dipping formations, and anticlines)
formation pressures normal for the deepest part of the zone will be transmitted to the
shallower end, where they will cause abnormal pressure conditions. However, the pore fluid
density will act against the normal hydrostatic pressure at the base of the reservoir, and so the
pressure at the top of the reservoir will equal the normal hydrostatic pressure at the
hydrocarbon/water contact minus the hydrostatic pressure of the fluid column.

In the presence of hydrocarbons in anticlines, overpressures are encountered in the potential


pay section, whereas normal hydrostatic pressure conditions may still exist at and below the
oil/water contact (Fertl 1976).

Figure 1-1: Hydrocarbon Density Effect

Phenomena of Overpressure 1-1 Ver. 1.0


Halliburton Overpressure

In Figure 1-1 above, pressure just inside the reservoir at 5000 feet can be calculated as shown
below.

g = 0.14 psi/ft
Pn = 0.45 psi/ft

At the gas/water contact at 5500 feet, pressures are normal for the area, ie

5500 x 0.45 = 2475 psi


At D1, theoretical Pn is
5000 x 0.45 = 2250 psi

The true pressure within the reservoir top is given by

Po = D2 x Pn - [(D2 - D1) x g ]

= 5500 x 0.45 - [(5500 - 5000) x 0.14]


=2405 psi
Therefore

2405
= 9.25 ppg EMW
5000 x 0.052

Consequently, when drilling through a reservoir or thick sandstone sequence, if the pressure
is known at the top then it is possible to calculate the pressure at any point through the
section.

Figure 1-2: Calculation of Pressure at Base of Reservoir


Using Figure 1-2 above, pressure may be calculated as shown below.

PB = PT + HYD

PT is obtained from RFT, DST, or kick data as available. If, in the example, PT = 12.5 ppg
EMW, then
PT = 12.5 x 2000 x 0.052 = 1300 psi

Phenomena of Overpressure 1-2 Ver. 1.0


Halliburton Overpressure

Hydrostatic of the water column in the reservoir section is

8.88 x 500 x 0.052 = 216.6 psi


Therefore
PB = PT + HYD = 1300 + 216.6 = 1516.6 psi = 11.66 ppg EMW

Clearly pressure at the base of the sand is less than that at the top.

In a hydrocarbon reservoir with gas, oil and water the pressure at each of the contacts can be
calculated by inputting the respective densities of fluids into the above equation.

Figure 1-3: Pressure Profile Through Reservoir

Norwegian Sector
Frigg = 443 feet
Dutch Sector
K7 = 135 feet
K8 = 114 feet
K14 = 656 feet
Leeuwarden = 92-98 feet
British Sector
Leman = 800 feet (max)
Indefatigable = 50-419 feet
Viking = 330-450 feet

Table 1-1: Example Gas Column Heights in North Sea

Phenomena of Overpressure 1-3 Ver. 1.0


Halliburton Overpressure

1.1.2 Undercompaction (Sedimentary Loading)


Rapid deposition of sediments (with respect to geological time) can mean that fluid flow has
been so restricted that it has not yet escaped. The central North Sea contains some highly
overpressured Tertiary shale sequences, which have been deposited to a depth of some 11000
ft (3350m) in about 60m years. However, actual sedimentation rates have been produced by
Donato and Tully (1981) showing maximum deposition of some 5 cm/1000 years. The
dominant formation for this area is a soft grey clay (gumbo). Due to both rapid sedimentation
and low permeability, the pore fluid in this clay has not yet escaped and has given rise to
overpressured shales (maximum pressures in the Oligocene often correspond to higher
deposition rates). Given greater geological time it is probable that compaction will continue
and fluid will eventually be squeezed out producing a normally compacted sequence.

Carstens (1978) and others have noted how thin (1-3m) limestone bands can act as seals
preventing fluid expulsion in these Tertiary clays. In addition, the clays also display
abnormally high porosity resulting in lowered densities, low sonic velocities, and low
electrical resistivities. Often the limestone cap rocks act as perfect seals.

The greatest difference between shales which have a perfect seal and those which are rapidly
deposited, and therefore have an imperfect seal, is the increase in pressure upon entering
them. The imperfectly sealed formation is characterised by a gradual build-up of pressure
(over several metres to hundreds of metres) while the perfect seals display a rapid pressure
build-up as soon as the zone is penetrated.

Limestones capping the Kimmeridge clay are common in the North Sea, and often conceal
large pressure changes.

1.1.3 Aquathermal Pressuring


Work by Kennedy & Holser (1966) first indicated that water heated in a closed vessel will
increase about 125 psi/F. Thus a formation which is completely isolated can have pore
pressure increased by 1000 psi as a result of only 8 F increase in temperature.

In a typical sedimentary sequence, the geothermal gradient can be expected to range from 1.0
to 2.5 F/100 ft. Thus for an isolated formation fluid, pressure resulting from aquathermal
effects may range from 1.25 to 3.2 psi/ft. Magara (1975) used a figure of 1.4 psi/ft for the
Gulf Coast and showed that through aquathermal pressuring an overpressured sequence may
become equal to the overburden pressure. For example, a shale sequence with pore pressure
3600 psi becomes isolated at 8000 ft. If this formation was then to be buried to 20000 feet the
pore pressure would equal

3600 + (12000 x 1.4) = 20400 psi

ie approximating to total overburden pressure (assuming overburden gradient of 1 psi/ft).

Phenomena of Overpressure 1-4 Ver. 1.0


Halliburton Overpressure

Aquathermal pressuring could therefore be used to account for areas where the pore pressure
is greater or equal to overburden pressure.

1.1.4 Aquifers
In most instances an aquifer is a shallow sand which outcrops on nearby mountains at an
elevation appreciably higher than that of the well. Water entering at the outcrop influences the
pressure encountered in the well-bore. Although this pressure is essentially hydrostatic, it
gives the illusion of geopressure because of the increased column height.

Figure 1-4: Aquifer

1.1.5 Charged Upper Sands


High pressures can occur in shallow sands if they are charged by gas from lower formations.
This condition can also result from a poor surface casing cement job, casing leak, or a
blowout in a nearby well. Upper sands can also be highly pressured if gas developments are
trapped by very rapid deposition. This occurrence is relatively rare.

In the Middle East a similar type of phenomenon is shown by highly pressured shale sections
deposited within a massive salt section.

Incomplete sediment compaction refers to the rapid burial of low permeability sediments such
as clays or shales. At the time of deposition, such sediments can have a porosity in excess of
50% and are subsequently compacted following further deposition. Under this compaction the
porosity decreases and water that is expelled finds its way into other, more permeable,
sediments. If burial is rapid then there is little time for fluid movement and so the excess fluid
will help to support the overburden. Hence overpressuring of the excess pore fluids occurs.

Phenomena of Overpressure 1-5 Ver. 1.0


Halliburton Overpressure

1.1.6 Tectonic Loading


Tectonic loading occurs when pressure is increased due to earth movements. One of the most
common causes here is faulting. Thrust faulting also produces overpressures where the rock
rapidly moves over an as yet undisturbed sequence. This is similar to rapid loading as
described above.

Uplift

Formations at depth which are normally compacted may be uplifted to a shallower depth.
Should the original pressure be retained, abnormal pressure gradients will result. In the
example below the pressure gradient of the formation is seen to double.

Figure 1-5: Pressure Due to Uplift


The geological process which uplifts a buried formation also tends to lift the overburden. It
follows that uplift can only generate abnormal pressures when accompanied by another
geological process which reduces the relief between the buried rock and the surface. The
magnitude of pressure is therefore a function of the depth of burial and the degree of uplift.

For the same degree of uplift, higher pressure gradients result from shallower depths of burial.
Table 1-2 shows the effect of a 2000 foot uplift from various depths.

Original depth Uplifted depth Original New pressure Equivalent


(ft) (ft) pressure (psi) (psi/ft) Mud Weight
(ppg)
10000 8000 4650 0.58 11.2
8000 6000 3720 0.62 12.0
6000 4000 2790 0.70 13.5
4000 2000 1860 0.93 18.0

Table 1-2: Pressure Alteration Through Uplift

Phenomena of Overpressure 1-6 Ver. 1.0


Halliburton Overpressure

Faulting

In the event of subsurface movement creating severe faulting, deeper fluid pressures may
escape to shallower formations. This behaviour is similar to charged upper sands (1.1.5).
As in charged sands, abnormal pressures will persist if the higher pressure does not dissipate
to the surface.

Regional growth faults contribute to the origin of abnormal pressures by re-distributing


sediments and placing permeable zones opposite impermeable zones, thus inhibiting the flow
of fluids to regions of hydrostatic equilibrium. Faults may prevent the expulsion of water
during the compaction process whereupon the shales in such a zone remain of abnormally
high porosity.

Fracture zones may also allow the transmission of formation pressures upwards from a depth
at which they are normal to a shallower horizon where they represent abnormally high
pressure.

Figure 1-6: Faulting


1.1.7 Salt Diapirism
The upward movement of low density salt due to its buoyancy can disturb the normal layering
of sediments, producing pressure anomalies. Overpressured zones often occur due to the
faulting and folding actions associated with diapirism. Additionally, the salt may act as an
impermeable seal preventing lateral dewatering of clays.

Phenomena of Overpressure 1-7 Ver. 1.0


Halliburton Overpressure

Figure 1-7: Salt Diapirism

1.1.8 Mud Volcanoes


This mechanism, as with salt diapirism, refers to the upward movement of a low density
plastic zone, in this case shale. This is common in the Caribbean Sea, especially in Jamaica.

Figure 1-8: Mud Volcano

1.1.9 Clay Diagenesis

Phenomena of Overpressure 1-8 Ver. 1.0


Halliburton Overpressure

As montmorillonite alters to illite during clay diagenesis, inter-layer bound water is desorbed
and becomes free water. Large volumes of water are released by this process, which can
increase the pore fluid pressure.

Note that the increased pore fluid pressure will eventually act against the diagenetic process,
as the pressure causing expulsion of the fluid from the clay layers equilibrates with the pore
fluid pressure. As a consequence, diagenesis is halted by overpressures.

1.2 SUBNORMAL FORMATION PRESSURES


Subnormal formation pressures are those corresponding to a gradient less than hydrostatic.
The primary mechanisms causing development of underpressures are discussed below.

1.2.1 Artificial Production


Subnormal pressures are commonly created when hydrocarbons and/or water are produced.

Production (unless compensated for by a strong water drive) will reduce pore pressure and
therefore cause compaction. In turn, this may lead to land subsidence (eg Ekofisk).

Where freshwater aquifers have been tapped the reduction in hydrostatic head can cause
subnormal pressure. As an example, the Texas Panhandle has gradients ranging from 0.36 to
0.39 psi/ft due to this mechanism.

1.2.2 Precipitation
In arid areas such as the Middle East, the water table may be found hundreds of feet below the
surface. As a consequence of this underpressured formations may develop as the hydrostatic
gradient commences at the water table only, causing a subnormal gradient when measured
from the surface.

1.2.3 Potentiometric Surface


This mechanism relates to the structural relief of a formation, and can result in under- or over-
pressured reservoirs. There is a spontaneous electrical potential between formations which
indicates the flow of electrical current (water flows to cathode). This flow of current moves
fluids through the porous media.

1.2.4 Osmosis

Phenomena of Overpressure 1-9 Ver. 1.0


Halliburton Overpressure

Strong salinity contrasts in lenticular sand bodies are favourable to osmotic action which may
result in subnormal pressures. In the Morrow Sands (Oklahoma) there is a regional transition
from under- to over-pressures.

1.2.5 Temperature Change


If subsurface temperature is reduced the pore pressure must decrease, especially when gas is
present. As sediments and pore fluids are buried during deposition, temperatures rise and, if
allowed to expand, the fluid density decreases.

However, the magnitude of this effect is very small. With a thermal gradient of 1.5 F/100ft,
the gradient at 20000 feet would be 0.432 psi/ft compared to 0.442 psi/ft.

1.2.6 Epeirogenic Movements


Changes in elevation can cause abnormal pressures in some formations open to the surface
laterally but otherwise sealed. Thus if the outcrop is raised the formation pressure becomes
abnormally high and vice-versa.

Pressure changes are seldom caused by changes in elevation alone since associated erosion
and deposition are also significant factors. Loss or gain of water saturated sediments is also
important.

1.2.7 Formation Foreshortening


This mechanism may occur in areas of modern tectonic activity, for example along the flanks
of the Rocky Mountains. It is suggested that during a compression process upwarping of the
upper beds and downwarping of the lower beds can result. The intermediate beds must
expand to fill the voids left by this process. It is then possible for more competent,
intermediate beds to have a subnormal pressure gradient.

Figure 1-9: Formation Foreshortening

Phenomena of Overpressure 1-10 Ver. 1.0


Halliburton Overpressure

1.2.8 Decompressional Expansion


It has been observed that, in gas reservoirs in the Appalachian Region, subpressure occurred
in reservoirs associated with shales in areas which had been eroded. This erosion may have
been caused by the adsorption of water in clay minerals as the overburden pressure and
temperature decreased and pore volume increased due to expansion of the crystal structure.

1.2.9 Different Densities of Oil


Density variation causes underpressure due simply to the changes in hydrostatic pressure
generated by the variable density fluids. When measured against water pressures,
hydrocarbons must generate lower than water-normal.

1.3 PROBLEMS ARISING FROM OVERPRESSURE


There are several advantages to being able to detect and measure abnormal pressures. A
detailed study of local conditions before, during, and after drilling of the well allows for the
following.

a) More effective well programming.


b) Maximum ROP with minimum mud weight.
c) More economical selection of casing points.
d) Minimum trouble from lost circulation and kicks.
e) A better understanding of local geology and drilling problems.

Standard drilling practice dictates that mud weight used should be as close as possible to the
balance point with formation pore pressure. There are several reasons for this.

a) To minimise the risk of lost circulation.


b) To minimise the risk of differential sticking.
c) To minimise formation damage.
d) To maintain an optimum ROP.

1.3.1 Lost Circulation


This is one of the most common problems associated with rotary drilling. It has been defined
as the loss of drilling mud in quantity to the formation. It may occur at any depth where the
total pressure against the formation exceeds the pressure exerted by the formation.

Coarse permeable formations such as gravels may not be able to "take" a mud. This factor
depends on the ratio between the pore openings and the particle sizes found in the mud. Mud
loss only occurs to zones with relatively large openings. Mud losses to cavernous and/or
vuggy formations and sometimes to reefs, gravel, or other permeable zones are usually

Phenomena of Overpressure 1-11 Ver. 1.0


Halliburton Overpressure

predictable in a given area because they occur in definite formations which are easily
traceable.

However, mud losses may also occur in formations with no permeable zones or caverns. Such
losses occur in fissures or fractures, which may occur naturally, or may be created, enlarged,
or extended by mechanically imposed pressures. In many cases natural fractures are
impermeable under normal conditions, but when some critical pressure is reached or
exceeded, they open up and take mud. Once such a fracture opens up, the mud lost will tend
to wash out and enlarge the fracture. This is serious as later pressure reductions (reduced mud
weight) may not close the fracture and so the loss of mud will continue.

1.3.2 Differential Sticking


If a high degree of overbalance exists between the mud column and the pore pressure,
excessive filter cake build-up is likely to occur. In this situation differential sticking of pipe to
the borehole walls can be a problem, because with increased filter cake the area of contact
will be larger.

1.3.3 Formation Damage


Formation damage incurred when the mud weight is excessive includes such problems as
washouts, excessive borehole corrosion, reservoir flushing and contamination.

1.3.4 Maintaining Rate of Penetration


In experiments, Vidrine and Benit (1968) found that ROP can be reduced by up to 70% as
differential pressure was increased from 0 to 1000 psi. They found the sensitivity of ROP to
differential pressure was greatest when large diameter bits were used. Use of excessive
overbalance (over 1000 psi) means that changes in WOB, RPM and other factors do not alter
the ROP to any great degree.

Fontenot and Berry (1975) suggest that, given adequate cleaning, maximum penetration rate
should occur at zero differential pressure. A possible exception would be drilling of very
weak formations where a negative differential pressure could cause spilling of rock into the
hole.

Depth (ft) Mud Weight Phyd Formation Formation Overbalance


(ppg) Gradient Pressure (psi)
(ppg) (psi)
1000 12.0 624 10.0 520 104
2000 12.0 1248 10.0 1040 208

Phenomena of Overpressure 1-12 Ver. 1.0


Halliburton Overpressure

5000 12.0 3120 10.0 2600 520


10000 12.0 6240 10.0 5200 1040
15000 12.0 9360 10.0 7800 1560

Table 1-3: Overbalance Increase With Depth

Phenomena of Overpressure 1-13 Ver. 1.0


Halliburton Overpressure

1.4 SUMMARY OF OVERPRESSURE IN THE NORTH SEA


Abnormal pressure regimes occur in all geological formations to some extent and their
magnitude varies widely.

Tertiary sediments are mainly clays and shales and may be overpressured for much of their
thickness. Pressures of around 0.52 psi/ft are often encountered at a depth of 4500 feet, and
pressures up to 0.8 psi/ft have occurred locally. The uppermost sediments of the Tertiary have
a small quantity of expandable (high montmorillonite and illite content) clay due to the
material source of the clay deposited after the Late Miocene. In the Forties area the
Recent-Pliocene sediments are normally pressured. The major source of sediments for the
North Sea were the large European rivers which derived their Pliocene-Mid Miocene clays
from pre-existing shales in Scandinavia, the British Isles, and NW Europe in general. These
are completely dewatered and lithified, and also have a large content of quartz. Drilling this
section involves minimum problems and can usually be carried out using a normally weighted
uninhibited mud system.

From 4000 to 7500 feet there is an expandable clay associated with the North Sea Tertiary
volcanic episode. This lasted from the Palaeocene to early Mid Miocene. The clays are largely
of volcanic origin and are rich in montmorillonite and illite. The clay is young, has a low
density, and is in the first stage of dewatering. The decrease in density from the upper clays
would normally reflect an apparent abnormal pressure or undercompaction, but in this case
the formation is still undergoing compaction and is consequently quite plastic and prone to
hydration. Mud weights, however, are commonly raised to improve hole stability. The
proportion of expandable clay continues to increase to a depth of 9500 feet in the Lower
Eocene. In the Forties area, the Miocene, Oligocene, and Eocene rocks are overpressured
before the normally pressured Palaeocene is reached. In the Ekofisk area the Tertiary clays
may require a mud weight of 0.62 psi/ft to keep the well-bore open. This is about equal to the
overburden gradient. In addition, these "gumbo" shales can cause other drilling problems such
as bit balling, blocked flowline, mud rings, and high viscosity mud. Below the gumbo shales
the pressure may stay high or reduce slightly.

In Cretaceous sediments, overpressured oil occurs in the Danian chalk formation of the
Ekofisk area.

In the Mesozoic marl-claystone-shale sequence of the Central Graben, overpressures of up to


0.9 psi/ft have been recorded. In the Jurassic of the Viking Graben, overpressures up to 0.69
psi/ft occur which may be transmitted into the rocks overlying the Jurassic. Whittaker and
Shaw (1975) suggest the following causes of overpressure in the Northern North Sea Jurassic.

Phenomena of Overpressure 1-14 Ver. 1.0


Halliburton Overpressure

a) Subcompaction.
b) Faulting as a cutoff or pressure escape route.
c) Hydrostatic communication by way of complex fracturing systems.
d) Loss of porosity as a result of crystallisation.
e) Diagenesis of montmorillonite.

In Triassic sediments overpressures have been found in gas bearing zones of the Bunter
Sandstone in the Southern North Sea area. Zechstein salts in the Leman Field contain
overpressured carbonates and sandstones. Examples of this include the porous sections of the
Platten- and Haupt-dolomite sandwiched between thick salt layers. Where the Zechstein salt
is thick and continuous a mud weight of 0.73 psi/ft is required to keep the hole open. Below,
the Zechstein appears in domes, and the pressure problems become extreme. The Zechstein is
a mixed salt with a high percentage of water of crystallisation. It therefore reacts very
plastically compared to sodium (NaCl) salts and pressure rises rapidly when entering the salt.
In the shales underneath the edges of the dome, the mud weight approaches overburden value
and remains high as deep as the section has been penetrated.

Phenomena of Overpressure 1-15 Ver. 1.0


Halliburton Overpressure

1.5 REFERENCES
Carstens H. 1978, Origin of Abnormal Formation Pressures in Central North Sea Lower
Tertiary Clastics. The Log Analyst, Vol 19 No 2 pp 24-28.

Donato J.A. & Tully M.C. 1981, A Regional Interpretation of North Sea Gravity Data,
Petroleum Geology of the Continental Shelf Of North West Europe. Heyden & Son Inc.

Fertl W.H. 1976, Abnormal Formation Pressures. Elsevier NY

Fertl W.H. & Chilingarian G.V. 1977, Importance of Abnormal Formation Pressures.
JPT, Vol 29 pp 347-354.

Fontenot J.E. & Berry L.N. 1975, Study Compares Drilling Rate Based Pressure
Prediction Methods. Oil & Gas Journal, Vol 73 No 37 pp 123-138.

Kennedy G.C. & Holser W.T. 1966 Pressure-Volume-Temperature and Phase Relations of
Water and Carbon Dioxide. Geol.Soc.Am.Mem.97.

Magara K. 1975, Importance of Aquathermal Pressuring Effect in Gulf Coast. AAPG


Bulletin Vol 59 No 10 pp 2037-2045.

Vidrine D.J. & Benit E.J. 1968, Field Verification to the Effect of Differential Pressure
on Drilling Rate. JPT, July 1968 pp 676-682.

Whittaker A.H. & Shaw S. 1975, Problems Encountered in the Delineation of Formation
Pressure Regimes in the Jurassic of the Northern North Sea. JNNSS
Vol 14 No 1.

Phenomena of Overpressure 1-16 Ver. 1.0


Halliburton Overpressure

Phenomena of Overpressure 1-17 Ver. 1.0


Halliburton Overpressure

SECTION 2
OVERBURDEN GRADIENTS

2.1 TERZAGHI AND PECK EQUATION


Terzaghi and Peck ( 1948 ) developed the following relationship

S=+P

where S = overburden pressure


P = pore pressure
= formation matrix stress

This equation forms the basis of all pressure engineering.

Under normal circumstances, as compaction of sediments occurs due to the weight of


overlying sediments, porosity is reduced and the pore fluids squeezed out so that the weight
of the overburden is supported by the rock matrix alone. If, however, the pore fluid is unable
to escape then compaction is unable to continue in the normal fashion. A proportion of the
overburden must therefore be supported by the pore fluid.

In the former case above, the pore fluid pressure will simply be the hydrostatic pressure of the
fluid (the normal pressure). In the latter case the pore fluid pressure will be greater than the
normal pressure, ie overpressured.

Figure 2-1: Subsurface Pressure Trends

Overburden Gradients 2-1 Ver. 1.0


Halliburton Overpressure

2.2 OVERBURDEN PRESSURE


Overburden pressure is essentially the cumulative weight effect of the overlying sediments
above any point in the well.

Compaction of shales with depth will result in an increase in their bulk density per unit of
rock due to the reduction in pore space.

If bulk density (b) and porosity () are plotted against depth, the trends below are typically
obtained. The relationship between the two trends is apparent; each varying exponentially
with depth and inversely with each other. If these trends are plotted on semi-log paper,
straight lines are obtained. These are known as normal compaction trends (NCTs).

Figure 2-2: Log Response in a Normally Compacted Sequence

Any shale points falling on these lines will be normally pressured. Deviations from these
lines are caused by changes in the pore pressure gradient, or by lithological variations.

2.2.1 Overburden Stress


Overburden stress is strictly a function of the average bulk density of overlying rocks

S = b g D

where S = overburden stress


b = average bulk density
g = gravitational acceleration
D = depth

Overburden Gradients 2-2 Ver. 1.0


Halliburton Overpressure

Figure 2-3: Offshore Overburden Gradients

Overburden Gradients 2-3 Ver. 1.0


Halliburton Overpressure

As bulk density (especially in shales) is a function of increasing depth due to compaction, an


exponential relationship is found. The resulting increase of S with depth plots as a curve.

Formation densities can range from between 2.2 to 2.5 SG in shales. Thus, for a depth of
10000 feet the overburden pressure ranges from

9500 psi < S < 11000 psi

Dickinson (1953) states that a good average value of 2.3 SG would give an overburden
gradient of 1.0 psi/ft. Note that this value has been used by other authors as a definitive value,
although it should only be used where actual data cannot be obtained.

Donato and Tully (1980) have quoted some average North Sea density figures.

Eocene 2.44 gm/cc


Palaeocene 2.21 gm/cc
Cretaceous 2.44 gm/cc
Jurassic/Triassic 2.43 gm/cc

2.2.2 Water Depth and Air Gap


It is important to recognise that on offshore locations the weight of water acting upon the
sediments will exert an overburden stress which must be accounted for in any calculations.

Also, as all depths are measured from RKB and all pressure readings are with reference to the
hydrostatic column of drilling fluid, the air gap must also be included in calculations.

2.3 DETERMINATION OF OVERBURDEN STRESS GRADIENT


Initially the value of overburden gradient was taken to be S/D = 1.0 psi/ft

At any depth
S/D = sum of rock matrix + pore fluid above that depth

that is
S = 0.433 (1 - ) b + 0.433 ( ) n

where = porosity (assumed fluid filled)


b = average density of sediments above depth of interest (SG)
n = density gradient of fluid (SG)
0.433 = conversion constant (SG to psi/ft)

For example

Overburden Gradients 2-4 Ver. 1.0


Halliburton Overpressure

= 12% (0.12)
b = 2.5 SG (20.8 ppg)
n = 1.08 SG (0.465 psi/ft)

S =[0.433(1 - 0.12) x 2.5] + [0.433(0.12) x 1.08]


= 1.0087 psi/ft.

If the matrix density = 2.3 SG (19.0 ppg) and the porosity is zero, then

S = 0.052 x 19.0 = 0.988 psi/ft.

However, true overburden gradient is defined by

0.433
S/D = b (z) dz
D

where D = vertical depth (ft)


b(z) = bulk density varying with depth
0.433 = conversion constant

In practice this is approximated to


S/D = 0.433 b
instantaneous S/D in psi/ft

Therefore summing all the instantaneous overburden gradients gives

0.433 x b x DI
S/D =
D

where DI = depth interval


D = total depth

For example, if water depth = 200 ft and water density = 1.07 gm/cc

instantaneous S/D = 0.433 x 1.07


= 0.463 psi/ft

If the first formation increment = 2.0 gm/cc, then

S = 0.433 x 2 x 200
= 173.2 psi

Overburden Gradients 2-5 Ver. 1.0


Halliburton Overpressure

92.6 + 173.2
S/D = = 0.655 p.s.i./ft
200 + 200

Note for offshore overburden gradients the air gap must be included

Continuing the example above, if the air gap is 60 ft

0 + 92.6 + 173.2
S/D = = 0.57 p.s.i./ft
60 + 200 + 200

In the tabulated example below (figure 2-4), a constant depth interval (50m) is chosen which
allows the need to consider depth in the overburden gradient calculation to be eliminated, as
the depth values on both top and bottom of the equation cancel out.

Therefore all that is required using the constant depth interval method is to

a) average the density value over the interval (column 2)


b) apply the conversion constant to convert from g/cc to psi/ft (column 3)
c) sum the values in column 3 (column 4)
d) divide by the number of intervals (column 5) to obtain the variable overburden
gradient in psi/ft.

Note that using this method will generate figures in psi/ft irrespective of whether the depth
interval chosen is in feet or metres.

Overburden Gradients 2-6 Ver. 1.0


Halliburton Overpressure

Figure 2-5: Comparison of Overburden Gradients from Land and


Offshore Wells

Overburden Gradients 2-7 Ver. 1.0


Halliburton Overpressure

Alternatively, the variable overburden gradient down to the top of the first formation
increment can be calculated in one step.

For example
air gap = 50 ft
water depth = 150 ft
First formation density = 2.03 SG at 400 ft

Assuming that the formation density from sea bed to 400 ft is constant at 2.03 SG, then the
variable S/D at 400 is

(50 x 0 x 0.433) + (150 x 1.07 x 0.433) + (200 x 2.03 x 0.433)


S/D =
(50 + 150 + 200)
(0 + 69.5 + 175.8)
= = 0.613 p.s.i / ft
400

From this point real data from FDC or sonic log may be used to continue evaluation of
overburden gradient as in the table shown in figure 2-4 above.

2.4 AGIP FORMULA FOR CONVERSION OF SONIC TO BULK


DENSITY
Generally, direct density reading wireline logs (FDC type) are only run in reservoir sections
due to their high cost and limited application (ie to the oil company). Therefore, for most of
the well (and certainly for top-hole sections) no direct density data will be available.

It is possible however to obtain estimated density values from the t sonic log data by
application of the Agip formulae.

For Cemented and Compacted Formations

( t - t mx )
= (1)
( t f - t mx )

For Uncompacted Sands and Shales

1.228 ( t - t mx )
= (2)
( t + t f )

where
= porosity

Overburden Gradients 2-8 Ver. 1.0


Halliburton Overpressure

t = formation transit time (sec/ft)


tf = fluid transit time (sec/ft)
tmx = matrix transit time (sec/ft)

Porosity is related to density by the following

b = m (1 - ) + f (3)

where
b = formation density
m = matrix density
f = pore fluid density

Combining equations 1, 2 and 3, and using

average matrix density = 2.75 gm/cc


fluid density = 1.03 gm/cc
average matrix travel time = 47 sec/ft
average fluid travel time = 200 sec/ft

then for compacted and consolidated formations

t
b = 3.28 - ( )
89

and for uncemented formations

( t - 47)
b = 2.75 - [2.11 ]
( t + 200)

Both equations give similar results for compacted formations (40 - 60 sec/ft) but are quite
different for longer transit times. The second equation satisfies the density evaluation for
most types of formations. However, massive carbonate and evaporite sequences represent an
abnormal situation since the AGIP formula calculation represents a density value which is
much too high. One must therefore use an estimated density value for these formations (table
2-1 below).

Density values can now be processed to compute the overburden pressure gradient (figure 2-
6).

Overburden Gradients 2-9 Ver. 1.0


Halliburton Overpressure

Formation Actual Density t to input


Anhydrite 2.96 25
Halite 2.17 140
Polyhalite 2.78 44

Table 2-1: Sonic Values for Input to Computer

REFERENCES
Addis M.A. & Jones M.E. 1985 Volume Changes During Diagenesis. Marine &
Petroleum Geology, Vol 2 pp 241-246.

Bellotti P. & Giacca D. 1978 AGIP Deep Drilling Technology Pt 3, Pressure Evaluation
Improves Drilling Programs. Oil & Gas Journal, September 1978, Vol
76 No 37 pp 76-85.

Dickinson G. 1953 Geological Aspects of Abnormal Reservoir Pressures in


Gulf Coast Louisiana, AAPG Bulletin, Vol 37 No 2 pp 410-432.

Donato J.A. & Tully M.C. 1981 A Regional Interpretation of North Sea Gravity Data,
Petroleum Geology of the Continental Shelf of North West Europe.
Heyden & Son Inc.

Eaton B.A. 1972 A Theory On The Effect of Overburden Stress on


Geopressure Prediction from Well Logs. JPT Aug, pp 929-934.

Jones M.E. & Addis M.A. 1985 On Changes in Porosity and Volume During Burial of
Argillaceous Sediments. Marine & Petroleum Geology Aug, pp
247-253.

Lang W.H. (Jr) 1980 Determination of Prior Depth of Burial Using Interval Transit
Time. Oil & Gas Journal, Vol 78 No 4 pp 222-232.

Terzaghi K. & Peck R.P. 1948 Soil Mechanics in Engineering Practice. John Wiley &
Sons, N.Y., pp 729.

Overburden Gradients 2-10 Ver. 1.0


Halliburton Overpressure

Figure 2-6: Overburden Gradient Worksheet

Overburden Gradients 2-11 Ver. 1.0


Halliburton Overpressure

Figure 2-7: Estimated Formation Density From Sonic Log

Overburden Gradients 2-12 Ver. 1.0


Halliburton Overpressure

SECTION 3
dc EXPONENT

3.1 DEVELOPMENT OF THE DRILLING RATE EQUATION


Drilling rate is a function of the parameters listed below.

(a) Rotary Speed (RPM) (f) Overburden pressure


(b) Weight on Bit (WOB) (g) Pore pressure
(c) Tooth efficiency (h) Drillstring effects
(d) Drilling hydraulics (i) Matrix strength
(e) Differential pressure (P) (j) Lithology variations

Jorden and Shirley (1966) stated that, with constant drilling parameters in uniform lithology,
ROP should decrease exponentially with depth as compaction increased. Therefore, to
normalise drilling rate variations in these parameters should be allowed for, because under
normal drilling conditions it is often impossible or inadvisable to maintain them as constants.

The effects of these parameters on drilling rate are discussed below.

3.1.1 Rotary Speed


Wardlaw (1968) proposed that when all the mechanical energy applied to the bit was used in
rock fracture, and none in rock removal ("perfect cleaning"), drilling speed (R) would be
directly proportional to rotary speed (N), ie

RN

In practice, however, there is a non-linear relationship between ROP and rotary speed since
perfect cleaning does not exist due to overbalance (chip hold-down effect) and insufficient
circulation rates.

Thus
R Na
where a = rotary exponent

Vidrine & Benit (1968) determined a empirically from field data, and quote a value of
between 0.4 and 1.0.

Bourgoyne & Young (1974) report 0.4 (for very hard formations) to 0.9 (for very soft
formations).

dc Exponent 3-1 Ver. 1.0


Halliburton Overpressure

Figure 3-1: ROP v RPM as a Function of Hole Cleaning

Prentice (1980) proposes an exponential increase in penetration rate with increased RPM due
to "dwell time", ie the time available for the bit weight to be applied as each tooth strikes the
bottom of the hole. Thus, as RPM increases the bit weight is not as effectively applied to the
bottom of the hole, with a resultant decrease in ROP.

Prentice suggests a rotary exponent of 0.6 for continuous depositional basin drilling.

3.1.2 Weight On Bit


WOB is more correctly defined as force applied per unit area (ie the effective weight on bit
per unit area of bit cutting structure). This includes variations in bit size, tooth shape and
distribution, actual weight on bit (W), and threshold weight (Wo).

Threshold weight is defined as the actual weight at which the bit will commence to drill (ie
when the tooth penetration of the formation occurs).

Vidrine & Benit (1968), and Maurer (1962), state that under perfect bottom hole cleaning
conditions, drilling rate is proportional to the square of bit weight. This is supported by
experiments conducted by Somerton (1959).

Vidrine & Benit (1968) further suggest -

R (W-Wo)

Field tests in shale show that with an 8 1/2" bit Wo = 10 klbs, ie a weight to diameter ratio of
1180 lbs/in.

dc Exponent 3-2 Ver. 1.0


Halliburton Overpressure

Figure 3-2: Effect of WOB on ROP

Combs (1968) used


R Waw

but found that when aw = 1, acceptable results were obtained.

Bourgoyne & Young (1974) developed

W Wo a
R( - )
D D

where W = WOB
Wo = threshold weight
D = bit diameter
a = 0.6 to 2.0

Prentice (1980) states that ROP is directly related to bit weight between two limits. These are
The low (or threshold) weight which in very soft unconsolidated formations may have
a zero or even negative value. This suggests that the bit can achieve penetration by
jetting action alone.

The upper weight known as the "flounder point", which occurs whenever the bit teeth
are completely imbedded in the formation thus bringing the cone face into contact
with the bottom of the hole.

dc Exponent 3-3 Ver. 1.0


Halliburton Overpressure

3.1.3 Tooth Efficiency


Tooth efficiency takes into account three main effects

(i) Efficiency of the original cutting structure


(ii) The minimum effective cutting structure (ie the point of tooth wear at which the bit
will cease to drill)
(iii) The rate at which the bit will lose its efficiency

A dull bit can mask changes in the drilled formation. The effect of dulling is exaggerated for
longer tooth bits, and is hence related to bit type.

Vidrine & Benit state that the relationship of drilling rate with bit wear is not linear.

1
R
f(T)

where f(T) = function of tooth wear

f(T) can be approximated by


f(T) = (1 + 2.5T)

dc Exponent 3-4 Ver. 1.0


Halliburton Overpressure

T = normalised tooth wear


= 0 for a new bit, = 1 for a worn bit

The drilling rate is governed by the factor 2.5, and is dependant on both bit type and the
nature of the formation.

If the loss in drilling efficiency and tooth wear relationship is assumed to be linear, then

Ro = R1(1 + 2.5T)

where Ro = drilling rate with sharp bit (wear = 0)


R1 = drilling rate with dull bit (wear = 1)

Bourgoyne & Young (1974) assumed an exponential decrease in drilling rate with tooth wear.
R e-h

h = fractional tooth height worn away


or
R = ae-h

a = constant depending on bit type and formation

3.1.4 Drilling Hydraulics


Hydraulics depends on pump pressure, nozzle size and type, and mud rheology. If too little
hydraulic action is applied, insufficient bottom hole cleaning and reduced penetration rates
will result. However, hydraulic action in excess of that necessary for efficient bottom hole
cleaning may increase ROPs by the jetting action.

Combs (1968) suggests

Q
R( )
aq

3 Dh Dn

where Q = flow rate (gpm)


Dh = hole diameter (in)
Dn = nozzle diameter (in)
aq = hydraulics exponent
R = ROP (ft/hr)

Combs suggests a value of 0.3 for aq.

dc Exponent 3-5 Ver. 1.0


Halliburton Overpressure

The group

Q
Dh Dn

was selected as this controls the cross flow velocity beneath the bit, and hence hole cleaning.
The term actually represents the momentum flux or hydraulic impact per unit area of hole.

Wardlaw (1968) suggested that drilling rate is proportional to the square


of the jet velocity.

R x P d = KV 2

where R = ROP
Pd = differential pressure across nozzle
V = jet velocity
K = constant

Bourgoyne & Young (1974) assumed an exponential increase in drilling rate with Reynolds
number.

Q
R = e( a 1 + a 2 ( ))
350 d n

al + a2 = constants
= mud density (ppg)
Q = flow rate
= viscosity (cp)
dn = nozzle diameter (ins)
350 = units constant

P)
3.1.5 Differential Pressure (
This parameter represents the difference between the drilling fluid column hydrostatic
pressure and the formation pore pressure (ie degree of overbalance).

P affects drilling efficiency by controlling the rate at which cuttings are cleared from
bottom. A high (positive) differential pressure may well introduce a chip hold-down effect
whereby loosened cuttings are held to the bottom of the hole.

Cunningham & Eenink (1959) reported from their experiments that the drilling rate decreased
when mud column pressure exceeded formation pressure due primarily to the chip hold down
effect, and secondarily by localised compaction and strengthening of the rock.

dc Exponent 3-6 Ver. 1.0


Halliburton Overpressure

Vidrine & Benit stated from field evidence that a 70% reduction in drill rate may be observed
as the differential pressure increased from 0 to 1000 psi and that sensitivity to P changes
was greatest for larger sized bits. They also suggested that with differential pressures >1000
psi changes in WOB, RPM, and other factors do not noticeably alter ROP.

Wardlaw (1968)

a
P = - b
R

where a & b = constant for formation drilling conditions

Eckel (1967) reported that differential pressure is the only pressure parameter which
significantly affects drill rate.

Fontenot & Berry (1975) suggest that, given adequate cleaning, maximum penetration rate
should occur at zero differential pressure.

3.1.6 Overburden Pressure


Increased depth of burial results in increased compaction, and hence increased compressive
strength. This results in a slow decrease in bit performance with depth.

3.1.7 Pore Pressure


Bourgoyne & Young (1974) assume an exponential increase in drill rate with pore pressure
gradient.

R = e( a1 + a 3 D )0.69 ( g p - 9.0)

where a1 & a3 = constants


D = Depth (ft)
gp = formation fluid gradient (ppg)

dc Exponent 3-7 Ver. 1.0


Halliburton Overpressure

3.1.8 Drillstring Effects


The penetration rate may decrease if the rotary torque fluctuates causing erratic drilling
action. Thus hole deviation, radical changes in string stabilisation, and the strength and dip of
the formation may all effect the ROP.

3.1.9 Matrix Strength


Matrix strength is often referred to as the drillability of the formation, and has been
compared to the compressive strength of materials as measured in laboratories. Formation
compaction, porosity, and density will modify the matrix strength.

Bourgoyne & Young (1974) assume an exponential decrease in penetration rate with depth in
a normally compacted formation.

Zoeller (1970) defined formation drilling strength as the measurement of the formation
resistance to failure or chipping when a wedge shaped flat crested tooth is pressed into the
formation.

Figure 3-4: Rock Bits Drill By Impact Fracturing

dc Exponent 3-8 Ver. 1.0


Halliburton Overpressure

3.1.10 Lithology
Matrix strength varies with rock type, and so lithology changes may considerably affect ROP.

Figure 3-5: Effect of Lithology on Penetration Rate

3.2 DRILLING RATE EQUATIONS

3.2.1 Bingham
Bingham proposed a generalised drilling rate equation to interrelate all relevant drilling
parameters (1964).

W d
R = a Ne ( )
D

where R = ROP (ft/hr)


N = rotary speed (rev/sec)
e = rotary speed exponent
W = WOB (lbs)
D = bit size (ft)

dc Exponent 3-9 Ver. 1.0


Halliburton Overpressure

a = matrix strength
d = formation drillability exponent

Bingham absorbed changes in P in the constant a.

3.2.2 Jorden and Shirley (1966)


Jorden and Shirley solved Bingham's equation for d (the drillability exponent)

R
log10 ( )
d = 60N
12W
log10 ( 6 )
10 D

where R = ROP (ft/hr)


N = RPM (rev/min)
W = WOB (lbs)
D = bit size (ins)

They assumed e = 1 (linear increase in ROP with RPM) and a = 1 (constant lithology,
removing the need to derive a matrix strength constant). Hence d exponent becomes lithology
dependant.

Other modifications were to make the equation compatible with oilfield units,ie

60xN converts revs/min to revs/sec


D/12 converts bit size inches to feet
106 allows simple scale for d exponent

When (R/60N)< 1, then (R/60N) varies inversely with ROP, and so

d Exponent varies inversely with ROP

When drilling constant lithology, d exponent will increase with depth, compaction, and P

d exponent is not compensated for mud weight, SPP (ie hydraulics), and bit wear. These
variables should therefore be kept as constant as possible.

dc Exponent 3-10 Ver. 1.0


Halliburton Overpressure

3.2.3 Rehm and McClendon (1971)


This paper proposed the corrected d exponent to account for changes in mud weight, hence
the term dc Exponent.

MW 1
dc = xd
MW 2

where dc = modified d exponent


MW1 = normal pressure gradient
MW2 = mud weight (preferably ECD)
in other terms

Pn
dc = d x
ECD

3.3 QUANTITATIVE PRESSURE EVALUATION FROM THE dc


EXPONENT

3.3.1 Ratio Method


The relationship between the d exponent and its controlling parameters is exponential. The
normal compaction trend (NCT) may be expressed as

log dc = KD + K'

where D = depth
K = gradient of linear trend (function of normal pore pressure gradient)
K'= intercept at zero depth (function of matrix strength)

Deviation of dc exponent values from NCT on semi-log graph paper can be related to
overpressure magnitude by

dco
Po = Pn x
dcn

where Po = actual pore fluid pressure


Pn = normal pore pressure
dco = observed dc exponent
dcn = dc exponent from normal trend

dc Exponent 3-11 Ver. 1.0


Halliburton Overpressure

By rearranging the above formula an overlay can be produced (cf Eaton overlay below).

The ratio method should not be used in this over-simplified state as it ignores the effect of the
variable overburden gradient in controlling compaction trends. Although this effect is
reflected in the dc exponent trend, it is not accurately defined by it.

3.3.2 Equivalent Depth Method (Matrix Stress Equation)

Figure 3-6: Equivalent Depth Method for Pressure Evaluation

This method relies on the assumption that the matrix stress () is equal where tool response is
the same, since the dc exponent varies directly with compaction and formation drillability.
The technique is based on the relationship described by Terzaghi and Peck (1948).
S=+P

In figure 3-6 above, the pore pressure at 5000 feet is required.This is referred to as the depth
of interest (Di). However, at Di the value of is unknown.

dc Exponent 3-12 Ver. 1.0


Halliburton Overpressure

A line is constructed through the point of interest at Di and projected vertically upwards to
intersect any other point with the same dc exponent value, ultimately crossing the Normal
Compaction Trend at the equivalent depth (De). At De both the pore pressure (P) and
Overburden Gradient (S) are known, therefore solve for . As the value of must be the
same at Di (equal tool responses) this figure may be substituted into the equation

P=S-

and the equation solved for pore pressure.

This technique can also be expressed by the equation

( S i x Di ) - (( S e - Pe ) x De )
Pi =
Di

where Pi = pore pressure at depth of interest (psi/ft)


Pe = pore pressure at equivalent depth (psi/ft)
Si = overburden at depth of interest (psi/ft)
Se = overburden at equivalent depth (psi/ft)
Di = depth of interest (ft)
De = equivalent depth (ft)

As the component of rock matrix in the bulk rock increases due to compaction, resulting in
reduced porosity, the formation becomes increasingly harder to drill. An overpressured
formation which has, by definition, increased porosity for it's depth will also have a reduced
matrix stress component for that depth.

The equivalent depth method, as its name suggests, attempts to relate these values to the
depth where they would be normal.

The major flaw in the theory occurs when the equivalent depth of a particular overpressured
formation is found to be several hundred feet above the rig floor, such as in the case of large
scale overpressures developed at relatively shallow depths.

3.3.3 Eaton Equation (1975)


Po S S P dc
= - [( - n ) ( o )1.2 ]
D D D D dcn

The Eaton equation produces the most accurate estimation of formation pressures based on dc
exponent techniques.

dc Exponent 3-13 Ver. 1.0


Halliburton Overpressure

Eaton quotes 0.5 ppg accuracy for use of his equations worldwide (assuming b = 1.2 for dc
exponent).

The Eaton equation is derived from Terzaghi & Peck (1948). Since is difficult to define in
the field, the equation is re-arranged as shown below.

S=+P
=S-P
P=S-
Therefore
P = S - (S - P)

Eaton introduced the ratio

dco 1.2
( )
dcn

in order to quantify the effect of overpressure on the drilling rate, giving

dco 1.2
P0 = S - [(S - Pn ) ( ) ]
dcn

The b exponent of 1.2 may vary from region to region.

3.3.4 Determination of b Exponent


If an accurate indication of pore pressure from RFT, DST, or kick data is available, then the b
exponent can be back-calculated and used in the Eaton equation to give a greater degree of
accuracy (up to 0.2 ppg)

dco b
Po = S - [(S - Pn ) ( )] (1)
dcn

Rearranging (1) above gives

dco b
(S - Pn ) ( ) = (S - Po )
dcn

therefore

dco b S - Po
( ) =
dcn S - Pn

dc Exponent 3-14 Ver. 1.0


Halliburton Overpressure

Taking logs

dco S - Po
b log ( ) = log ( )
dcn S - Pn

When rearranged, gives

S - Po
log ( )
S - Pn
b =
dc
log ( o )
dcn

3.3.5 Back-Calculation of dcn


Using similar logic to 3.3.4 above, the NCT can also be back-calculated from RFTs, well
tests, etc.

Equation for dc normal

dco
dcn = S - Po 0.833
( )
S - Pn

The normal trend can now be pivoted through this calculated point to fall on the best fit line
through the dc exponent data.

Two such points may be joined together to establish a more accurate line. If several calculated
points are available, then a best fit line through these should be constructed, allowing for the
relative accuracy of the source of pore pressure, ie kicks and DSTs are most accurate,
followed by RFTs, then pressures extrapolated from swabbed gasses, eg connection gas.

The calculated NCT is superior to any created by simple visual inspection of dc exponent
data, as the dc data are affected by criteria outwith their definition, eg bit wear.

3.3.6 Construction of Eaton Overlay


The Eaton overlay uses Eaton's equation defined above to construct a series of lines
representing the value of dc exponent to be expected at a particular depth for a series of
different pore pressures.

dc Exponent 3-15 Ver. 1.0


Halliburton Overpressure

Figure 3-7: Construction of an Eaton Overlay

dc Exponent 3-16 Ver. 1.0


Halliburton Overpressure

The advantage of this method is that the overlay can be extrapolated to TD of the well as soon
as a reasonable NCT has been constructed. Thus, as actual dc exponents are obtained, they
may be plotted onto the overlay plot and the pressure read by considering the position of each
dco in relation to the overlay lines.

Additionally, there is no need to repeat the Eaton calculation at each depth point where a dc
exponent value has been calculated.

The procedure for construction is as shown below.

1. Draw NCT through good shale points, or points calculated as in 3.3.5 above.

2. Rearrange the Eaton equation to obtain that below.

S - Po 0.833
dco = ( ) x dcn
S - Pn

This equation solves for expected dc exponent values at known values of pore
pressure.

3. A grid can now be constructed by inputting values for Po, eg 9, 10, 11, and 12 ppg at
several different depths. The points for each pressure are then joined together to
establish the overlay grid lines.

4. Overpressure values can be quickly read off with reference to these lines of equivalent
mud weight.

3.4 INTERPRETATION OF THE dc EXPONENT


The dc exponent is a valid indicator of overpressures in impermeable but porous formations
which are undergoing increasing compaction with depth. The formations where the dc
exponent is applicable include

Shales
Claystones
Argillaceous Siltstones
Calcareous Claystones

3.4.1 Shales/Claystones

dc Exponent 3-17 Ver. 1.0


Halliburton Overpressure

These are the rock types which give the most accurate trends for quantitative overpressure
evaluation. The presence or absence of fissility within the shales has no effect on the dc
exponent result. However, accessory components may influence the result. These include

a) Small amounts of sand (as sandstone laminations) which will increase the average
drilling rate and therefore move the dc trend unpredictably. Dispersed sand, which
has no grain-to-grain contact and has no intergranular porosity, will not affect the dc
exponent.

Any other constituents which are dispersed will not affect the trend eg pyrite,
glauconite, and mica.

b) Anhydrite and pyrite will alter the trend if they are laminar.

Detailed analysis of the cuttings will help to determine the "true" shale points. If there are any
variations in dc exponent then the cuttings should be examined.

3.4.2 Argillaceous Siltstones


If these are not grain-supported they will behave exactly like shales.

If they are grain-to-grain supported then the trend will shift to the left or right depending on
cementation. Care must be taken at this point, as some trends will remain parallel and can
therefore be shifted, while others will not be usable.

3.4.3 Calcareous Claystones


Carbonate deposits as thin limestone beds will disrupt the trend. These stringers may also act
as cap rocks for a potentially overpressured zone below.

Large amounts of carbonates as concretions or fossils will not affect the trend unless they are
grain supported with inter-granular porosity.

Calcareous claystones will affect the trend line. These will tend to move the dc exponent to
the right, ie increase it's value. Dc exponent should remain constant if the amount of carbonate
is constant, but as this is hardly ever calculated accurately it is often impossible to determine.

The main danger of these calcareous claystones is that they can push the trend line across to
the right, giving the impression of lower overpressures than are actually present.

Therefore the dc exponent must be interpreted with reference to calcimetry results.

dc Exponent 3-18 Ver. 1.0


Halliburton Overpressure

3.4.4 Shifting NCT


The dc exponent plot often shows marked shifts through a variety of reasons including

a) Bit dulling (teeth and bearings)


b) New bit type (insert, diamond, turbine, etc)
c) Major change in a drilling parameter (not always fully compensated for by the
equation)
d) Bit hydraulics
e) Unconformities

The dc equation attempts to compensate for changes in hole size, mud weight, and drilling
parameters. However, the equation is not perfect and cannot accommodate large changes in
any parameter. In fact there is no compensation made at all for bit type, bit hydraulics and bit
wear (all assumed to optimum).

As a result of this, it may sometimes be necessary to shift the trend. This should only be done
after much interpretation of other data. These realignments should only be done for large
changes.

Procedure for Shifting NCT


A change should be made by the following method

1) Realign the first two singles drilled on a trend/trend basis. Do not align end to end as
the start and end points are often spurious.

2) Move the rest of the points by the appropriate factor. Do this by multiplying the dc
exponent by the factor

dc1
dc2

3) Do not shift the trends for coring, as ROP is generally variable.

4) Do not shift for hydraulics in top hole, again the actual trend is variable.

5) Bit dulling is generally considered to take place quickly at the end of the bit's life,
usually the last few metres.

Once the new bit is drilling, check that the new trend continues along the previous
trend.

dc Exponent 3-19 Ver. 1.0


Halliburton Overpressure

In rare cases, very rapid dulling will take place (indicating wrong bit selection). The
real trend will be established only in the first few metres drilled.

6) At casing points use the wireline log information to redefine true pore pressure values
obtained during drilling.

Figure 3-8: Effect of Bit Wear On ROP

Bit Shifts
Bit shifts are often misinterpreted and may lead to erroneous pore pressure values being
quoted. It is therefore generally preferable to set a NCT from the raw dc exponent plot with no
shifts applied.

Changes in bit size at casing points generally shift the dc exponent trend-line to the left as a
result of faster ROPs due to drilling a smaller diameter hole.

The dc exponent data may be normalised to a standard hole size by using the formula below to
standardise to 12 1/4".
3
Hole size
Hole size (norm) =
12 1 / 42

Ice Sheet Compaction

dc Exponent 3-20 Ver. 1.0


Halliburton Overpressure

The phenomenon of ice sheet compaction (fig 3.9) is often encountered when drilling in high
latitudes. Good NCTs are apparent in top hole sections, due to the increased competence of
the near-surface sediments caused by the weight of a once-present ice sheet.

As a result of this, NCTs shifted too far to the right are often set. The influence of the ice
sheet weight is often dissipated after a few hundred feet, with the result that the dc exponent
appears to cut back from the normal indicating an increase in pore pressure.

A similar phenomenon has also been observed on t sonic plots.

Figure 3-9: Overcompaction Caused by Ice Sheet Loading

Unconformities
Unconformities represent a "special case" in the interpretation of the dc exponent as they often
change the character of the normal trend line.

Setting of trend lines must, for the sake of accuracy, be confined to shale formations of the
same provenance and compaction history. Upon crossing an unconformity, entirely different
sedimentary conditions may be encountered resulting in not only a shifted normal trend line,
but also a different slope. It is important therefore to be aware of such phenomena either
through prior knowledge or good geological control.

dc Exponent 3-21 Ver. 1.0


Halliburton Overpressure

Ideally, a new normal trend line should be established upon crossing an unconformity.

dc Exponent 3-22 Ver. 1.0


Halliburton Overpressure

3.5 NOMOGRAPH FOR QUICK dc EXPONENT CALCULATION

Figure 3-10: Nomograph for Calculating dc Exponent


1. Enter the nomograph on line (1) and project a line through the appropriate RPM value
on line (2) onto line (3).

2. Enter the WOB line (7) and project a line through (6) (Bit size) onto line (5).

3. Join the resultant points of lines (3) and (5) and read off the value for d exponent
where it intersects line (4).

4. Correct the d exponent for mud weight by application of the formula

Pn
d c exponent = d x
ECD

dc Exponent 3-23 Ver. 1.0


Halliburton Overpressure

REFERENCES
Allen J.H. 1977 Optimising Penetration Rate Pt 1, Determining Parameters that Affect
Rate of Penetration. Oil & Gas Journal Vol 75 No 41 pp 94-107.

Bingham M.G. 1965 A New Approach to Interpreting Rock Drillability. The Petroleum
Publishing Company.

Black A.D., Dearing H.L. & Di Bona B.G. 1985 Effects of Pore Pressure and Mud
Filtration on Drilling Rates in a Permeable Sandstone. JPT, Vol 37 No
10 pp 1671-1681.

Black A.D., Tibbitts G.A., Sandstrom J.L. & Di Bona B.G. 1985 Effects of Size on Three
Cone Bit Performance in Laboratory Drilled Shale. JPT, Vol 37 No 9
pp 473-481.

Borel W.J. and Lewis R.L. 1969 Ways to Detect Abnormal Formation Pressures. Pet.
Eng. Vol 41 No 10 pp 101-109.

Bourgoyne A.T. and Young F.S. 1974 A Multiple Regression Approach to Optimal
Drilling and Abnormal Pressure Detection, Aug, pp 371-384.

Combs G.D. 1968 Prediction of Pore Pressure from Penetration Rate. SPE 2162, AIME
43rd Annual Fall Mtg Houston.

Cunningham R.A. & Eenink J.G. 1959 Laboratory Study of the Effect of Overburden,
Formation and Mud Column Pressure on Drilling Rates of Permeable Formations. Trans
AIME Vol 216 pp 9-17.

Eaton B.A. 1975 The Equation for Geopressure Prediction from Well Logs. 50th Annual
SPE of AIME Fall Mtg Reprint No SPE 5544 11pp.

Eaton B.A. 1976 Graphical Method Predicts Geopressures Worldwide. World Oil, Vol
182 No 6 pp 51-56.

Eckel J.R. 1958 Effect of Pressure on Rock Drillability, Trans AIYE Vol 213 pp l-6,
also JPT Apr 1967.

Fertl W.H. and Timko D.J. 1973 How Downhole Temperatures, Pressures Affect
Drilling Pt 9, Novel Ways to Detect Abnormal Pressures. World Oil,
Vol 176 No 2 pp 49-50.

Fontenot J.E. & Berry L.N. 1975 Study Compares Drilling Rate Based Pressure
Prediction Methods. Oil & Gas Journal Vol 73 No 37 pp 123-138.

dc Exponent 3-24 Ver. 1.0


Halliburton Overpressure

Hawkes S.L. 1985 How to Analyze Bit Records to Increase Penetration Rates. Petroleum
Engineer International Vol 57 No 5 pp 72-84.

Jorden J.R. & Shirley O.J. 1966 Application of Drilling and Performance Data to
Overpressure Detection, JPT Vol 18 No 11 pp 1387-1394.

Maurer W.C. 1962 The Perfect Cleaning Theory of Rotary Drilling. AIME Pet. Trans. Vol
225 p 1271.

Maurer W.C. 1966 How Bottom Hole Pressure Affects Penetration Rate. Oil & Gas
Journal, Jan, pp 61-65.

Moore P.L. 1982 How to Predict Pore Pressure. Pet. Eng. Inst. Vol 54 No 3, pp 144-152.

Prentice C.M. 1980 Normalized Penetration Rate Predicts Formation Pressures. Oil & Gas
Journal, Vol 78 No 32 pp 103-106.

Rehm W.A. and McClendon R 1971 Measurements of Formation Pressure from Drilling
Data. SPE 3601.

Somerton W.H. 1959 A Laboratory Study of Rock Breakage by Rotary Drilling. Petroleum
Trans AIME Vol 216 pp 92-97.

Vidrine D.J. & Benit E.J. 1968 Field Verification of the Effect of Differential Pressure
on Drilling Rate. JPT Jul, pp 676-682.

Wardlaw W.W.R. 1969 Drilling Performance Optimisation and Identification of


Overpressure. SPE 2388.

Zoeller W.A. 1970 The Drilling Porosity Log. 4th Annual Fall Mtg. SPE 3066.

dc Exponent 3-25 Ver. 1.0


Halliburton Overpressure

SECTION 4
OTHER INDICATORS OF ABNORMAL PRESSURE

4.1 GAS LEVELS


Gas enters the mud system as the formation is drilled by the bit. If a differential pressure is
created on bottom by a combination of low density mud and higher formation pressure, small
amounts of gas will enter the wellbore and increase the amount of background gas in the mud.
Background gas normally increases gradually in a transition zone.

Figure 4-1: Some Sources of Gas While Drilling (after Fertl)

Other Pressure Indicators 4-1 Ver. 1.0


Halliburton Overpressure

Pixler (1945) recommended the use of gas measurements for the detection of overpressures
and for warnings of impending blowouts. Goldsmith (1972) states that most impermeable
shales contain some gas, while abnormally pressured shales often contain large quantities of
gas. Fertl (1973) explains this by saying that comparatively free gas diffusion is possible
through clay as a function of the median pore size of clays or silty clays and the varying
diameter of gas molecules. Since overpressured shales have high porosity, diffusion will be
enhanced resulting in shale gas to be found over long impermeable shale sections. Low
salinity and high pressures increase the amount of solution gas in formation waters.

As these shales are circulated up the hole, gas pressure in the cutting pores explodes the shale
particles, releasing gas into the mud. This shale gas is the prime source of gas-cutting of the
mud, since gas may enter the mud from only two basic sources. These are

1 Gas flow due to underbalance

2 Gas evolving from drilled cuttings

Increasing the mud weight will only decrease this second effect indirectly, by reducing the
amount of shale sloughing.

Goldsmith states that gas flow into the wellbore through low formation permeabilities can
also occur. In such a situation, the degree of gas cutting can roughly be correlated with the
amount of underbalance.

4.1.1 Connection Gas


This is gas produced when circulation is stopped for short periods of time, eg to make a
connection. When the pumps are shut off, the total fluid pressure will decrease from dynamic
to static mud column pressure. This decrease in pressure, coupled with a further loss due to
swabbing when the bit is raised off the bottom, may allow some hydrocarbons to seep into
the mud, thereby increasing the concentration above the expected drilled gas level.
Connection gas peaks are generally short and sharp, although this depends on the bottoms up
time since gas will migrate upwards through the mud, and so tend to spread the peak. The
longer the circulation takes, the more spread due to migration will occur.

Correlation of gas levels with changes in mud weight can give an accurate indication of
formation pressure. For example, if a small mud weight increase suddenly decreases high
background gas levels and associated connection gas peaks, then it is reasonable to assume
that the formation pressure is only slightly below that of the new ECD.

Other Pressure Indicators 4-2 Ver. 1.0


Halliburton Overpressure

Figure 4-2: Gas Levels as Indicator of Pore Pressure

Referring to figure 4-2, it is noted that there is a reduction in bottom hole pressure at each
connection due to the swabbing effect of the pipe. When the pore pressure exceeds dynamic
mud column pressure connection gas will appear as peaks of produced gas. The connection
gas peak size, with respect to background gas levels, will increase as the pressure differential
increases. When the pore pressure finally exceeds dynamic mud column pressure, total
background gas readings will also begin to increase as the formation is now underbalanced
and is producing.

4.1.2 Trip Gas


This is produced by much the same mechanism as connection gas, although in this case the
swabbing effect caused by pipe movement is generally more sustained as stands of pipe are
pulled from the well. The width of a trip gas peak can give an indication of conditions at the
bottom of the hole. An early peak may indicate that swabbing has taken place some way up
the hole, usually due to insufficient formation of mud cake. Allowance must be made for gas
migration rates, which may range from 1000 to 2000 ft/hr, tending to spread the swabbed gas
considerably.

Other Pressure Indicators 4-3 Ver. 1.0


Halliburton Overpressure

This poor build up of cake can indicate that the pressure differential between wellbore and
formation has not allowed filtration and hence cake build up take place. Therefore, the early
onset of trip gas can indirectly relate to the state of balance in much of the open hole.

4.1.3 Other Factors Affecting Gas Readings


If pressure reduction from minor gas cutting is not significant, there are other aspects which
are quite dangerous. Recycling of the cut mud through the pumps decreases their efficiency.
Enough gas will render the pumps to be ineffective. The recycled gas plus the greater relative
influx from a lower pumping rate may cause a blowout.

Also, it should be remembered that gas-cutting can be caused by cavings. Many shales
contain enough gas to cause continued gas cutting long after they have been drilled. When
these shales wash out the mud will be gas-cut. This often results in false concern.

Other factors that may affect gas readings are

Pay Zones High gas concentrations in potential pay zones may diffuse into shales
located immediately above.

Kelly Air While making a connection air may be caught in the kelly. When this
air reaches the bit (and especially if the drill collars are large with
respect to the hole size) there may be a slight loss of hydrostatic head,
and more gas may be evolved from the formation. The gas will tend to
diffuse from the mud or formation into the air bubbles by density
segregation, thereby concentrating gas levels within that interval. This
probably has a greater affect than most hydrostatic pressure changes.

Downtime While tripping there may be inflow of gas due to the swabbing
mechanism discussed above, plus a continuous inflow from the more
permeable formations.

H2S and CO2 These gases have a thermodynamic behaviour different from that of CH4. Both
are soluble in mud (particularly oil-based muds). Expansion of H2S and
CO2 only takes place at low pressures high in the borehole (bubble
point), hence there is very little warning of kicks resulting from these
gases.

Degradation of Mud Additives Gases such as H2S and CO2 may originate from mud
additives that degrade due to high temperature. Lignosulphonates and other organic additives
can degrade at temperatures above 400 F.

Lignite Zones These are often associated with high gas readings.

Other Pressure Indicators 4-4 Ver. 1.0


Halliburton Overpressure

Diagenesis of Volcanic Ash Over geological time, diagenesis of volcanic ash results in three
components, namely clay minerals, CH4, and CO2. Drilling associated
shales causes gas cutting without directly reflecting formation pressure
variations.

Shale Diapirs These can expel large amounts of solids, fluids, and gases causing frequent gas
cutting but not necessarily overpressure.

Faults These often channel gas, causing localised gas flow into wells, and thus the
mud may become gas-cut.

Additives to Mud The addition of diesel or crude oil, and carbide for lag time
determination.

Thermodynamic Processes Clays exhibiting catalytic activity are present in both formation
and drilling fluid. Field tests show that hydrocarbon gases also
originate as a result of the grinding action of the bit creating a
temperature increase in the presence of a catalyst (clay mineral) and
organic matter in the rock (Fertl 1976).

4.2 INCREASED DRAG AND TORQUE


Drag (overpull) is the excess hook load over the free handling load. This excess in load may
be caused by bit balling, dog legs, deviated holes, differential sticking, extra cuttings coming
into the wellbore when drilling transition zones, the hole coming in around the drill collars
and bit, or insufficient cleaning of cuttings from the hole whilst drilling. Drag is noticed at
connections or when tripping, but may also result from keyseating, junk in the hole and exotic
bottom hole assemblies.

Undercompacted shales are considered to be plastic in nature. When a differential pressure


into the wellbore exists, these shales will tend to reduce the hole diameter to a value smaller
than the bit size. Increased torque should be noticed when underbalanced conditions prevail.
This is especially true if full gauge stabilisers are present in the drill string. Torque is useful
in detecting large increases in pore pressure, for example by crossing a fault line. Increased
torque as a result of underbalanced conditions is virtually unseen with differentials into the
bore hole less than about 1.0 ppg EMW. Torque can also be produced by an increase in the
size and amount of cuttings coming into the wellbore. A drastic increase in torque may mean
a locked cone on the bit, hanging up of full gauge stabilisers on limestone stringers, or a
change in pore pressure.

Torque and drag are not valid indicators when drilling high-angle directional holes.

McClendon (1977) states that torque will tend to increase in the transition zone with a low
density mud because a larger amount of cuttings will enter the borehole. Shale can tend to

Other Pressure Indicators 4-5 Ver. 1.0


Halliburton Overpressure

stick to and/or impede bit rotation, and bit teeth will take larger bites of the formation as they
are rotated. If the mud has been weighted to result in a hydrostatic pressure greater than the
formation pressure, the torque will be masked. In the presence of negative differential
pressure (underbalance) in the borehole, overpressured shales will tend to flow or heave into
the wellbore. Pilkington and McKee (1974) state that overpressured shales will tend to
slough when drilled underbalanced by 0.5 to 1.0 ppg thus causing the torque increase.

4.3 MUD PUMP PRESSURE


By observation of pump pressure one can observe u-tube effects in the borehole caused by
entry of less dense fluid into the annulus. This entry will reduce the hydrostatic pressure of
the annulus. As a result, heavier uncontaminated mud in the drill pipe will have a tendency to
u-tube down the drill pipe and up into the annulus, provided that the entry is slow enough to
allow the mud inside the drill pipe to maintain continuity with the mud inside the annulus.
The influx causing a reduced hydrostatic head will therefore cause a decrease in pump
pressure.

However, if the underbalance and rock permeability are large enough, an increase in drill pipe
pressure may be observed rather than the expected decrease, as the exit of the mud to the
annulus via the bit is opposed by the invading fluid.

4.4 PIT LEVELS


Any subsurface addition of fluid to the mud system will be indicated by a gain in pit level. If
a gain is recorded it is wise to stop drilling and to check for flow in the annulus.
Losses in pit level may be due to lost circulation.

4.5 FLOW METERS


An increased rate of flow returning from the annulus caused by fluid entering the wellbore
from the formation will be noticed before the corresponding rise in pit level, due to the
relative positions on the flowline of the flow sensor compared with the active pit sensor.
Rises in pit level are thus directly related to increased return of flow.

4.6 CUTTINGS
Though claystones and shales may have a good porosity, their permeabilities are generally
low. The net result may be an inward movement of the hole wall resulting in sloughing in
brittle formations and the production of sticky horizons in a gummy formation. Increased
ROP when entering transition zones will also cause an increase in the volume of cuttings

Other Pressure Indicators 4-6 Ver. 1.0


Halliburton Overpressure

observed at the shale shakers. Sloughing is encouraged by the presence of shale gas. An
increase in cuttings is generally dependant upon three factors

1 The amount of hole drilled below the point of balance between the hydrostatic mud
pressure and the pore pressure of the shale.

2 Increased drilling rate.

3 The magnitude of pressure differential into the wellbore.

If underbalanced conditions do exist then a mass of cuttings will be seen at the shale shakers.
These sloughed shale particles can be identified as long slivers of shale which are also larger
than regular shale chips seen in the overbalanced section.

4.6.1 Abnormal Trip Fill Up


The annulus should always be kept full of mud when pulling out of the hole. Mud must be
pumped in to compensate for the removed volume of drill pipe, otherwise hydrostatic head
may be sufficiently reduced to allow entry of formation fluids into the wellbore. Swabbing
during a trip aggravates this condition.

Thus if the volume of mud added is appreciably less than the volume of steel removed, then
the well may be flowing, especially if the annulus is observed to be full.

Increased fill bottom encountered after running into the hole may necessitate reaming to
bottom. This is generally a good indicator of the formation pressure being very near to the
hydrostatic head. Tight spots on trips can also be indications of the hole coming in.

4.6.2 Shale Density


As with dc exponent discussed above, establishment of a normal trend line is a critical factor.
Good lithological descriptions are necessary in order to make connections between
plot-points of constant lithotype, ie clean shale displays a lower density than a limy shale.
Experience has occasionally shown that depositional environment may affect shale densities,
ie reversals may be experienced on the transition from one formation to another
for example Eocene to Palaeocene. Thus if the geology of an area is well known, it is best to
attempt to establish a normal trend-line within a particular formational unit.

Quantitative techniques have been developed to aid continuous formation pore pressure
evaluation from shale density plots (Boatman 1967), but these are not widely used.

Shale density can be expressed as

Other Pressure Indicators 4-7 Ver. 1.0


Halliburton Overpressure

b = ( sh x w ) + (1 - ) g

where b = bulk density


w = water density
g = grain density (shale density)
= porosity as unit of 1

g - sh - b
% = = ma x 100
g - w ma - w

ma = matrix density
sh = shale density

Methods for Wellsite Evaluation of Shale Density


1 Mud Balance Method (Bulk Density)

Using mud balance, place cuttings in the cup to balance at the density of fresh water, 8.33
ppg (1 SG).

Fill the cup plus cuttings with water, and re-balance. The reading obtained at balance point is
W2.

8.33 1
Form. dens (g / cm3 ) = =
16.66 - W 2 2 - W2

The second equation is used where W2 is read in SG rather than ppg. The final value for both
equations is SG.

If oil-based mud is used and cuttings are washed with diesel, then diesel should be used as the
balancing fluid. The equation must therefore be modified to accommodate the density of
diesel (generally around 7.0 ppg) replacing that of water.

This method has the advantage of being fast and simple to perform, and uses a good quantity
of cuttings. Unlike other methods it does not require the selection of individual cuttings.

2 Density Column Method (Shale Density)

Zinc Bromide (ZnBr2) and water are partially mixed in a cylinder, and beads of known
specific gravity are suspended within the column. The column is graduated in density terms.

Other Pressure Indicators 4-8 Ver. 1.0


Halliburton Overpressure

Pieces of shale are lightly dried with filter paper and dropped into the column, floating at a
level of comparable density.

Zinc Bromide Mixing Instructions

This procedure will allow the creation of 1 litre of fluid. Ensure that the relevant page on the
manual Controls on Substances Hazardous to Health is read and understood before
proceeding.
a) Pour 500 cm3 of drinking water into large clean container, and add a few drops of
10% HCl.

b) Gradually add 5 bottles of zinc bromide crystals, stirring continuously (some heat will
be given off - this is normal).

c) After 5 bottles have been added the solution is saturated with density 2.6 g/cm3.
Crystals will be seen on the bottom of the container. Allow to stand for a few hours
and decant the solution into clean glass bottles.

The column is then mixed using 50:50 quantities of the heavy solution and clean water in a
graduated cylinder.

Figure 4-3: Use of Variable Density Column

Other Pressure Indicators 4-9 Ver. 1.0


Halliburton Overpressure

The graduated solution will not degrade in sunlight although after a time a small degree of
diffusion may occur, therefore keep out of direct sunlight. Agitation will cause mixing of the
column.

A normal compaction trend can be established by plotting shale density against depth. A
departure from this trend may indicate abnormal pressure.

There are several disadvantages to the method, which must be carefully considered.

a) Cuttings must to be circulated to surface before measurements can be made.

b) A number of readings must be taken to allow for sampling error.

c) The density of shale is decreased by exposure to water-based muds.

d) Small amounts of sand and secondary minerals within the shale cause error in the
measurements.

e) Shale gas can reduce the density of the shale.

4.6.3 Percent Montmorillonite (Shale Factor)


Montmorillonite clays are more porous and less permeable than most other clays. An increase
in montmorillonite clays will mean a lower density shale which may lead to misinterpretation
as an undercompacted shale. Several authors have associated montmorillonite with deep
marine shales in which sands are absent. Thus water cannot escape by way of the sands. They
speculate that if montmorillonite is associated with deep marine shales, increases in
montmorillonite tend to be associated with abnormal pressures. This however need not be the
case.

More simply, as diagenesis proceeds montmorillonite clays are converted to illite clays plus
water. Hence montmorillonite content should decrease with depth. However, overpressured
zones are assumed to be sections in which normal diagenesis (for that depth) has not taken
place. This is because in zones of abnormal pressure the pore fluid bears a greater part of the
overburden stress and the rock matrix a lesser part. Hence, because clay diagenesis is, in part,
a pressure dependant process the montmorillonite/illite ratio in the formation will increase.

Bound inter-layer water is released into the available pore volume between clay mineral
grains. Where fluids have become trapped in the clays, the resultant overpressuring of the
pore fluids acts against the release of this inter-layer water. Conversion to illite is therefore
halted.

Other Pressure Indicators 4-10 Ver. 1.0


Halliburton Overpressure

The ratio of montmorillonite to illite in cuttings samples is measured as the cation exchange
capacity (CEC). Montmorillonite has a much greater CEC than illite. The CEC is expressed
in milli-equivalents per 100 g of sample, and is also termed the shale factor. In essence then,
the shale factor should gradually decrease with depth and show an increase in abnormally
pressured regions (Nevins and Weintritt 1967, and Gill 1972).

Texture CEC Water Wt % Clay Content Wt Density


Content Water % (SG)
Clay
soft 20-40 free & 25-70 montmorillonite and 20-30 1.2-1.5
bound illite
firm 10-20 bound 15-25 illite and mixed layer 20-30 1.5-2.2
montmor/illite
hard 3-10 bound 5-15 High illite, trace 20-30 2.2-2.5
montmorillonite
brittle 0-3 bound 2-5 illite, kaolin, chlorite 5-30 2.5-2.7
firm-hard 10-20 bound 2-10 illite and mixed layer 20-30 2.3-2.7
illite/montmor

Table 4-1: General Shale Classification (Mondshine 1969)

Shale Factor Determination


1 Dry the sample.

2 Weigh out 1.0 g of sample and crush to fine powder with a mortar and pestle or coffee
grinder.

3 Add sample to about 30 cm3 of distilled water.

4 Wash the suspension into a titration flask, and acidify with about 2.0 cm3 of 2N
sulphuric acid, to neutralize non-clay cations in the sample.

5 Titrate the acidified suspension against methylene blue solution (3.74 g/l) vigorously
swilling the flask after each measured addition of methylene blue solution. After each
addition, take one drop of solution from the flask (on a stirring rod) and test on a filter
paper for the end point.

Other Pressure Indicators 4-11 Ver. 1.0


Halliburton Overpressure

On the filter paper, each drop will produce a dark blue stained aggregate of clay
particles surrounded by a fluid ring. When the fluid ring changes from colourless to
very pale blue, the end point is reached.

100
Shale Factor (meq / 100g) = x vol of titrant ( cm3 ) x normalityof titrant
wt spl

Since normality of titrant = 0.01N

Shale factor = volume of titrant (cm3) x 100 x 0.01


Shale factor = volume of titrant (cm )

4.7 CHLORIDES
When using a fresh water based mud, salt water entry from the formation will cause an
increase in chloride content of the mud filtrate. This amount depends on the contrast between
chlorides in the mud and chlorides in the formation. Difficulties arise in that routine mud
checks usually do not show the subtle changes in chloride content of the filtrate caused by
formation fluids. Also, the resistivity and conductivity of the mud is dependant to a large
extent on its temperature (allowances are made for this in the Foxboro conductivity probe)
but generally an increase in mud conductivity (assuming a constant chloride content of the
mud due to surface additions) will indicate increased pore fluid within the drilled formation,
and hence increased formation pore pressure.

4.8 FLOWLINE TEMPERATURE


Fluids within a transition zone are immobile, and as such may have above normal
temperatures. As a result, the formation gains heat and this is transferred to the mud during
drilling operations. This indicator has many pitfalls since flowline temperature also depends
on

1 The rate of circulation.

2 The time elapsed since last trip.

3 The volume of the mud system.

4 The mud type.

5 The amount of metal in the drill string.

Other Pressure Indicators 4-12 Ver. 1.0


Halliburton Overpressure

Flowline temperature is the drilling parameter most affected by surface events and least
affected by the conditions in the wellbore. It is far easier to change the flowline temperature
by adding water in the pits, for example, than by a change in temperature downhole.

Transition shales exhibit a lower thermal conductivity because of their higher fluid content.
They are hotter than the normally pressured formations above, because they do not transmit
heat radiating from the earth's core as efficiently. Long marine risers act as efficient heat
exchangers and may cool mud to a point where the flowline temperature plot becomes invalid
as an overpressure indicator. Cap rocks usually have a greater thermal conductivity than shale
and a reduction in flowline temperature may be seen, especially if the cap rock is thick.

Figure 4-4: Temperature Response When Drilling Into Overpressured


Formation

A more meaningful measurement may be T (the difference between temperature in and out).
T will normally decrease with depth due to longer circulation times at lower rates of
penetration. An increase in T can thus indicate entry into a transition zone. T plots reduce
the effect of ambient temperature changes and mud system additions on temperature data.

The thermal conductivity of water is considerably less than that of most rock matrix materials
and, since a characteristic feature of overpressured shales is the possession of a higher than
normal water filled porosity, it follows that overpressured shales also exhibit a lower than
normal thermal conductivity and a consequently elevated geothermal gradient.

As an overpressured shale lying below a normally pressured shale formation is approached,


the geothermal gradient (and consequently the flowline temperature gradient) decreases as a

Other Pressure Indicators 4-13 Ver. 1.0


Halliburton Overpressure

result of disturbance of equitemperature lines. On transition into the overpressured zone,


increases in geothermal gradient are reflected in increased flowline temperature gradient.

Other Pressure Indicators 4-14 Ver. 1.0


Halliburton Overpressure

4.8.1 Geothermal Gradients


Temperature increases with depth, therefore heat flow is from the earth's core outwards.

T
Q =
D

where Q = heat flow


T/D = geothermal gradient
= thermal conductivity of rock

From experiment and observation, heat flow varies from 40 - 80 mW/m2. This gives rise to
the following

1 Crystalline rocks are better thermal conductors than sediments.

2 Shales are very poor conductors.

3 Evaporates are very good conductors.

4 Pore fluids are very poor conductors, ie porosity will greatly reduce thermal
conductivity.

Thus the geothermal gradient is inversely proportional to thermal conductivity. Therefore the
geothermal gradient of shale will be high. However, geothermal gradients of sand will vary
depending on porosity. In a typical rock sequence, the geothermal gradient is never constant
but instead reflects the varying thermal conductivity of the sedimentary column.

During drilling, the formation is normally cooled due to the lower temperature of mud, the
degree depending upon factors such as

1 Rate of penetration.

2 Initial mud temperature.

3 Circulation rate.

4 Thermal properties of the rocks.

From research it seems that about four days are required for mud temperature to reach
equilibrium. Therefore various methods are used from consecutive wireline logging runs to
attempt to evaluate formation temperature, such as the Horner Plot (4.8.3).

Empirically, an average geothermal gradient of between 1.0 - 2.5 F/100 ft (18 - 45 C/km)
may be expected over large intervals.

Other Pressure Indicators 4-15 Ver. 1.0


Halliburton Overpressure

An understanding of temperatures is not only useful for overpressure detection. It is also


useful for the liquid window concept of oil maturation, ie oil preservation and maturation is
restricted to between 150 F and 300 F (65 - 150 C) (Pusey 1973 and Gretener 1978).

4.8.2 End - to- End Plots


While drilling a well, good mud temperature trends may be obscured by numerous bit trips,
etc. In order to clarify the trend it may be advantageous to plot flowline temperature as an
end-to-end (trend-to-trend) plot by laterally shifting individual bit run data (fig 4.5).

Figure 4-5: Migrated Plot Aids Temperature Interpretation

Other Pressure Indicators 4-16 Ver. 1.0


Halliburton Overpressure

4.8.3 Horner Plots and Bottom Hole Temperature (BHT)


When running a suite of logs, the BHT is recorded each time a new tool is on bottom. These
are plotted against the log of dimensionless time.

Dimensionless time is the ratio of time since last circulation stopped (t) to the sum of the
actual circulation time (T) and (t).

t
log dimensionless time = log ( )
t + T

Note that in this case, t is not related to T discussed in 4.8 above.

A straight line is drawn through these points and extrapolated to meet the x-axis, where

t
log( ) = 1
t + T

ie where t is a very large number relative to T. This gives the best available estimate of true
BHT.

The theory is based on the principle of thermal recovery, ie if a column of mud is left
undisturbed in the hole for an indefinite period it will eventually warm up to the same
temperature as the surrounding sediments. Thus temperature is plotted against the log of time
to give a straight line relationship. This allows projection of temperatures to some time in the
future and estimation of the final temperature of the mud, ie the estimated BHT.

Wiper trips and circulation will interrupt the thermal recovery time as cool mud from the
surface is pumped around the well. Therefore for subsequent logging runs a new Horner Plot
must be drawn up.

Other Pressure Indicators 4-17 Ver. 1.0


Halliburton Overpressure

Figure 4-6: Use of Horner Plot to Determine True BHT

Other Pressure Indicators 4-18 Ver. 1.0


Halliburton Overpressure

REFERENCES
Boatman W.A. 1967 Shale Density Key to Safer, Faster Drilling. World Oil Vol 165 Aug.

Fertl W.H. & Timko D.J. 1973 How Down Hole Temperatures, Pressures Affect
Drilling. Pt 9 Novel Ways to Detect Abnormal Pressures. World Oil Vol 176 No 2 pp 47-50.

Fertl W.H. 1976 Abnormal Formation Pressures. Elsevier, NY

Gill J.A. 1972 Shale Mineralogy and Overpressure: Some Case Histories of Pressure
Detection Worldwide Utilising Consistent Shale Mineralogy
Parameters. SPE of AIME Abnormal Subsurface Pressure Symposium,
Reprint No 3890, pp 121-136.

Goldsmith R.G. 1972 Why Gas Cut Mud is Not Always a Serious Problem. World Oil Vol
175 No 5, pp 51-54.

Gretener P.E. 1978 Pore Pressure: Fundamentals, General Ramifications and Implications
for Structural Geology. AAPG Continuing Education Course Note
Series No 4.

Kennedy G.C. & Holser W.T. 1966 Pressure-Volume-Temperature and Phase Relations of
Water and Carbon Dioxide, Geol.Soc.Am.Mem.97.

Lewis C.R. & Rose S.C. 1970 A Theory Relating High Temperatures and
Overpressures. JPT January 1970, pp 11-16.

McClendon R.T. 1977 Combinations of Drilling Data Pick Formation Pressures. Oil
& Gas Journal Vol 75 No 10 pp 102-110.

Mercer R.F. 1974 Liberated, Produced, Recycled or Contamination. 15th Annual


SPWLA Logging Symposium Trans. 20pp.

Mondshine T.C. 1969 New Technique Determining Oil-Mud Salinity Needs in Shale
Drilling. Oil & Gas Journal Vol 67 No 28.

Nevins M.J. and Weintritt D.J. 1967 Determination of Cation Exchange


Capacity by Methylene Blue Adsorption. Ceramic Bulletin, Vol 46 No
6.

Pham Thi Hang & Brindley G.W. 1970 Methylene Blue Absorption by Clay Minerals.
Determination of Surface Areas and Cation Exchange Capacities (Clay-Organic Studies
XVIII). Clays & Clay Minerals Vol 18.

Other Pressure Indicators 4-19 Ver. 1.0


Halliburton Overpressure

Pilkington P.E. & McKee R.E. 1974 Pressure Prediction and Detection Pt 2,
"Calculated" approach to pressures is preferred. Oil & Gas
Journal Vol 72 No 47 pp 129-131.

Pixler B.O. 1945 Some Recent Developments in Mud Analysis Logging. SPE 2026.
AIME

Pusey W.C. 1973 Paleotemperatures in the Gulf Coast using the ECR-Kerogen Method.
Trans Gulf Coast Assoc Geol Soc 23.

Timko D.J. & Fertl W.H. 1972 How Downhole Temperatures, Pressures Affect
Drilling, Pt 6, Correlating Geopressure Gradients with Hydrocarbon
Accumulations. World Oil Vol 175 No 6 pp 78-81.

Other Pressure Indicators 4-20 Ver. 1.0


Halliburton Overpressure

Other Pressure Indicators 4-21 Ver. 1.0


Halliburton Overpressure

SECTION 5
PRESSURE ESTIMATION FROM WIRELINE LOGS

5.1 EVALUATION OF PORE PRESSURE BEFORE DRILLING


There are four main methods by which prior indication of pressures likely to be encountered
while the drilling a well may be obtained.

5.1.1 Mud History and Drilling Reports


These are studies from offset wells in the area to be drilled. Mud weights will give a good
indication of both the location (depth) and magnitude of any abnormal pressures. Any
problems such as kicks, lost circulation, and differential sticking will also be found on mud
reports. However, in areas where swelling clays are encountered, mud weights as much as 2.0
ppg over the actual pore pressure may be required to control the formation. In this instance
mud weight is not an accurate indicator of pore pressure.

5.1.2 Geological Correlation


Pressured zones are often well delineated and thus can often be detected from correlation with
offset wells.

5.1.3 Geophysical Aspects


The value of geophysical data such as seismic and gravity surveys is rather ambiguous.
However, as indicated in fig 5-1, seismic data may be converted to interval transit time
(Pennebaker 1968), the resulting data being similar to an acoustic log. Attempts to use gravity
surveys to distinguish low density overpressured shale formations have met with little
success. However, shallow seismic techniques have located reworked glacial channels, mud
volcanoes, gas seeps, and gas sands. This type of interpretation should help eliminate shallow
blowouts, which are often the worst type of blowout.

Seismic two way travel times are recomputed to give interval transit times in a form
resembling an averaged acoustic log derived plot. The data are then interpreted as for a shale
sonic plot.

Wireline Logs 5-1 Ver. 1.0


Halliburton Overpressure

Figure 5-1: Comparison of Seismic Data With Sonic Log Data

5.1.4 Wireline Logs


Evaluation of wireline logs from offset wells is one of the most reliable methods of
estimating pore pressures prior to drilling a well. Most log interpretations are related either
directly or indirectly to porosity, and because sand porosity is so unpredictable, these
techniques are usually limited to shale sections. Shales, however, have the property of
compacting to a rather uniform and homogeneous structure. Thus pressures in the reservoir
may be estimated from the pressures of surrounding shales. Generally speaking porosity
decreases exponentially with depth, but under abnormal conditions water is not allowed to
escape and so porosity will no longer decrease. In most cases porosity will increase below the
top of the pressured zone.

Wireline Logs 5-2 Ver. 1.0


Halliburton Overpressure

5.2 GAMMA RAY LOG


To date, the gamma ray has been used as a log correlation and lithology indicating tool,
measuring the natural radioactivity of the formations adjacent to the borehole.

In nature there are three main radioactive elements. These are

Uranium, Thorium, and Potassium40

with K40 being the most abundant. All three elements are unstable and decay by emitting
particles of radiation (including gamma rays) until they eventually attain a stable atomic
structure.

Shales and clays generally have a high concentration of potassium, having been formed by
the decomposition of feldspars and micas which have a high K fraction. Clay particles also
absorb ions of heavy radioelements from mineralised waters during deposition. As a result,
shales and clays generally have high gamma ray counts.

Dark bituminous shales, eg the Kimmeridge Clay of the North Sea, often contain strong
traces of thorium and uranium and are very radioactive.

Sands formed by the mechanical erosion of quartz generally have low gamma ray values as
the stable crystal form of quartz precludes impurities such as radioelements. A badly sorted
and dirty sandstone, however, traps appreciable amounts of clay minerals which often cause a
reduction of the porosity and permeability of the sandstone as well as increasing its
radioactive response.

Carbonates generally have a low gamma ray response, although dolomitised limestones
may exhibit increased radioactivity due to the introduction of a low quantity of radioelements
by percolating waters.

However, other minerals can lead to misinterpretation, ie K-salts, arkosic sands, tuffs, and
glauconite. Gamma ray logs must always be used in conjunction with ROP and cuttings to
pick good shale points.

Gamma ray response is quoted in API units which relate to a standard permanent source in
the API test pit at the University of Houston, Texas. Gamma ray logs exhibit a degree of
statistical variation as the number of gamma rays reaching the tool varies with time due to the
random emission of radioactivity. Averaging circuits are placed in the tools to minimise these
statistical fluctuations. As a result, response times are increased so bed boundaries should be
picked at the point halfway between maximum and minimum deflections of the anomaly.

Zoeller (1983) proposes the use of natural gamma ray readings as a direct formation pressure
indicator.

Wireline Logs 5-3 Ver. 1.0


Halliburton Overpressure

In the Gulf Coast of the USA a progressive change in shale radioactivity can be related to
depth and rate of compaction. This is probably due to the de-watering of montmorillonite-rich
clays.

The deposition of Gulf Coast shales is associated with potassium-rich run-off water. As
compaction occurs, the pore volume is reduced with the consequent release of fluid into
highly permeable sand beds which are interbedded with the shales. However, the potassium
ions adsorbed onto the clay particles are not totally released during de-watering. Therefore as
compaction continues, the bulk volume decreases and the concentration of potassium ions
increases. This is reflected in a higher gamma ray intensity with depth. Obviously such a
relationship can only hold true as long as the depositional conditions remain constant, eg
source material, burial rate, etc.

Varying amounts of Uranium and Thorium could affect gamma ray responses, but reference
to a spectral gamma ray log for the well should indicate the limitations of the data.

If gamma ray response increases normally with depth, then departures from the normal
compaction trend may indicate changes in pore fluid pressure (increase in formation porosity
therefore relative decrease in K ion concentration, and consequently lower gamma ray
values).

While Zoeller's technique may be applicable to the Gulf Coast of the United States, with it's
thick Tertiary shale sequences of fairly constant provenance, it is extremely unlikely that the
same logic can be applied worldwide and certainly not to the North Sea which has undergone
a much more complex depositional history.

5.3 SONIC LOG


Primarily, sonic transit time may be considered as a function of lithology and porosity. If a
given lithology, such as a shale, is investigated then the sonic response will essentially be a
function of porosity variation. If sonic transit times of normally compacted shales are plotted
on a logarithmic scale against depth (natural scale), a linear trend will be exhibited, ie transit
time will decrease with depth. The fluid pressure exhibited with this normal compaction
trend will be hydrostatic. If overpressured formations are encountered, the data points will
diverge from the normal trend toward abnormally high transit times for any given depth. This
occurs due to the higher porosity of overpressured shales.

The interval transit time (t) is measured in microseconds per foot (s/ft).

A quick check to assess the validity of a sonic log involves noting the reading inside casing.
This should be 57 s/ft (187 s/m).

Wireline Logs 5-4 Ver. 1.0


Halliburton Overpressure

Wireline Logs 5-5 Ver. 1.0


Halliburton Overpressure

Figure 5-2: North Sea Sonic Log Response

Wireline Logs 5-6 Ver. 1.0


Halliburton Overpressure

Figure 5-3: Matrix Transit time Plot (after Hamouz and Mueller)

Hamouz and Mueller (1984) propose that a plot of

log(t - tma)

where tma = matrix travel time


t = observed sonic log value

is more valid than a simple log t plot, as the latter tends to have a curved trend, especially in
top-hole, whereas the former plots as a straight line. The authors assume a typical value for t
of 70 s/ft (figure 5-3).

5.3.1 Principle of the Sonic Tool


Pulses of sonic energy are transmitted to the formation and take several paths before reaching
the receivers (R1 and R2, figure 5-4). These are through the formation, through the mud, and
via the tool body. As the sonic waves are transmitted quickest through the formation, a timing
circuit is added to the system to cut out all late arriving waves.

Wireline Logs 5-7 Ver. 1.0


Halliburton Overpressure

Figure 5-4: Principle of the BHC Sonic Tool

The difference in time (t) for the same pulse to reach the first and second receivers is
recorded. This eliminates the effect of the mud and mud cake, leaving only the transit time of
the sonic wave through the interval of formation corresponding to the distance between R1
and R2.

These receivers are generally placed one to two feet apart.

5.3.2 Cycle Skipping


If the sonic signal is very weak then only the first receiver will be activated and R2 will wait
until it is triggered by a stronger signal from a subsequent cycle. This will cause excessively
high values of t, appearing as sharp deflections on the log.

Wireline Logs 5-8 Ver. 1.0


Halliburton Overpressure

Cycle skipping is often caused by vugular porosity, fractures in the formation, and also by gas
bubbles in the mud, all of which attenuate the sonic signal (figure 5-5).

Figure 5-5: Cycle Skipping as a Result of Fractures

5.4 RESISTIVITY

Wireline Logs 5-9 Ver. 1.0


Halliburton Overpressure

Resistivity is a measure of the ability of a formation to conduct an electric current. The solid
matrix is generally non-conductive, while the pore space may be filled either with
non-conductive hydrocarbons or by conductive saline water. Resistivity values are therefore
related to the amount and nature of the pore fluid and, ultimately, to the degree of porosity.

Resistivity measurements are influenced by such factors as proximity to adjacent formations,


the resistivity of the mud in the borehole, and the depth of invasion by mud filtrate. Tool
design attempts to minimise these effects.

Generally, combination resistivity tools such as ISF (Induction Spherically Focused) or DIL
(Dual Induction Laterolog) are run. These consist of three resistivity tools each with a
different depth of investigation (deep, medium and shallow) reading concurrently.

The deep reading tool will generally indicate true formation resistivity, while the shallow
readings will indicate the degree of invasion by mud filtrate and allow corrections to be made
accordingly. Shale resistivity should therefore be read from one of the deep indicators for the
purposes of pore pressure investigation.

Assuming homogeneous shale formation, the normal resistivity trend would be a gradual
increase with depth as compaction reduces porosity. Since overpressured zones contain more
pore fluid than would be normal for that depth, and the saline fluid is the conductive medium,
resistivity through an overpressured zone tends to diverge from the normal trend and show
as decreased resistivity.

Resistivity plots, although proposed by many American authors as indicating compaction of


sediments under normal and abnormal conditions, have been found to be of limited value in
the North Sea.

Resistivity by definition is critically dependant on the salinity of the pore fluid and therefore
demands a homogeneous source of both sedimentary material and of pore fluids. Any
variation in salinity, for whatever reason, will radically affect the resistivity tool response and
thus mask any variation due to pore pressure trends.

In the North Sea, salinity variations cause erratic tool responses, making it virtually
impossible to construct an accurate normal compaction trend through the data, especially in
top-hole sections. Overton and Timko (1969) propose that, as clays de-water, the conductivity
of the water changes. They believe the clay acts as an ionic membrane allowing only fresh
water to be desorbed until the shale becomes salt saturated. This results in a reversed trend.
Lithology variations within a shale sequence will also influence the resistivity data, often
making it impossible to use such plots.

5.5 DENSITY

Wireline Logs 5-10 Ver. 1.0


Halliburton Overpressure

Formation density can be read directly from the density log, and is plotted linearly on the
wireline log to give better separation between data points. A compensation curve is plotted on
the right of track 2; this value indicates the reliability of the readings and depends on caving
and mud cake thickness. Where this value exceeds two divisions the results tend to be wrong
and should be ignored. These values are automatically corrected onto the log. In heavy muds
with a high barite content, a negative correction is made, which makes the density
measurements unusable.

5.5.1 Formation Density Log (FDC)


The bulk density of normally pressured shales increases with compaction and therefore with
depth. The presence of undercompacted sediments is reflected as a reduction in bulk density.

Figure 5-6: Shale Density Response to Developing Overpressure

5.5.2 Density/Neutron (FDC/CNL)


Little divergence will be shown between the two trends in a clean, tight carbonate. Neutron
porosity in these formations is zero, and density around 2.70 g/cc.

In a gas-filled pore space, separation becomes extreme with the neutron tool moving to the
right indicating low porosity (often even negative).

The opposite is true for shales, which have a high hydrogen content. The neutron response
will move far to the left of the density.

Wireline Logs 5-11 Ver. 1.0


Halliburton Overpressure

This tool is a pad tool, and is therefore only accurate when the pad is in contact with the hole
wall. In badly washed out holes, readings may be unreliable; check with the calliper log to
assess hole geometry.

5.5.3 For All Logs


1 Beware of limestone stringers which may not necessarily display low gamma ray
readings. Their sonic, resistivity, and FDC measurements will all peak.

2 Ensure that each bed is thick enough for good readings to be developed.

3 Beware of picking sharp peaks on all logs, since the statistical nature of many of the
tools causes variation in tool response to be noted even in constant lithology. Instead,
pick averages over short intervals rather than sharp peaks.

Figure 5-7: Tool Response Related to Bed Thickness

5.5.4 Density/Sonic Crossplot


These can be used to evaluate the porosity, or water content of the clay fraction, in the
presence of a variable silt fraction.

A typical shale sand triangle is constructed using quartz, the dry clay mineral, and water as
the apices (figure 5-8).

Using standard log interpretation, V clay (clay fraction) and xp (crossplot porosity) lines
can be constructed. A new value, the clay porosity (or water content) can be estimated by
construction or computed from

Wireline Logs 5-12 Ver. 1.0


Halliburton Overpressure

xp
clay =
V cly + xp

Wireline Logs 5-13 Ver. 1.0


Halliburton Overpressure

Figure 5-8: Density-Sonic Crossplot for Porosity Determination

V clay thus derived represents the % of the clay fraction of the formation, provided no
effective porosity (which may occur in shaly sandstones) is present. A plot of % water (V
clay) on a log-scale against depth (natural scale) can then be made.

Wireline Logs 5-14 Ver. 1.0


Halliburton Overpressure

5.5.5 Using Data Quantitatively


Quantitative techniques proposed by Hottman and Johnson (1965) assume that a good deal of
regional data is available. They use empirical data from logs with well tests to develop a
correlation between pore pressure gradient and resistivity departure ratio.

R sh normal
R sh observed

Eaton (1972) developed an empirical relationship which works quite well where

R sh observed 1.5
P / D = S / D - [(S / D - Pn / D) ( ) ]
R sh normal

where P/D = pore pressure gradient (psi/ft)


S/D = overburden stress gradient (psi/ft)
Rsh = resistivity of shale (ohm/m)
Pn/D = normal pore pressure gradient (psi/ft)
R normal = normal trend extrapolated to depth of interest
R observed = observed resistivity at depth of interest

The above relationship is an empirical development from Terzaghi and Peck's (1948)
equation
S=+P

After further analysis (Eaton 1975), he revised the exponent value from 1.5 to 1.2 ie

R sh observed 1.2
P / D = S / D - [(S / D - Pn / D) ( ) ]
R sh normal

Eaton later developed this relationship for use with dc exponent (Section 3) and t sonic data.

For Sonic Data

tn 3
P / D = S / D - [(S / D - Pn / D) ( ) ]
to

Wireline Logs 5-15 Ver. 1.0


Halliburton Overpressure

5.6 FORMATION PRESSURE TESTS (RFT)


Most wireline logging companies have a pad tool which can measure directly formation
pressures. These must also measure depth and mud hydrostatic pressure. The tool records
formation and mud pressure values.

Interpretation is carried out by considering the following

a) If the recorded hydrostatic pressure values are consistent with the mud weights, then
the recorded formation pressures should also be accurate.

b) If hydrostatic pressures vary, then formation pressures will be suspect.

c) If hydrostatic pressures are consistent, but do not agree with the mud weight, then
suspect the mud balance is off calibration. This may be checked with water.

d) If hydrostatic pressure equals formation pressure, then the well is either on balance or
more likely the seal on the testing tool has failed.

Wireline Logs 5-16 Ver. 1.0


Halliburton Overpressure

REFERENCES
Cronkhite D.P. 1984 Calculating Porosity from Sonic and Bulk-Density Logs. Oil & Gas
Journal Vol 82 No 8 pp 70-71.

Dumont A.E. & Purdy V.S. 1976 Use of Seismic Data Can Cut Arctic Drilling Costs.
World Oil Vol 182 No 1 pp 71-74.

Eaton B.A. 1972 A Theory on the Effect of Overburden Stress on Geopressure


Prediction from Well Logs. SPE of AIME Abnormal Subsurface
Pressure Symposium, Reprint No. SPE 3719, pp 15-22.

Eaton B.A. 1975 The Equation for Geopressure Prediction from Well Logs. 50th Annual
SPE of AIME Fall Mtg Reprint No. SPE 5544, 11 pp.

Fertl W.H. 1974 Practical Formation Pressure Evaluation from Well Logs. Petrol Eng,
Vol 46 No 4 pp 56-70.

Fertl W.H. 1981 Open Hole Crossplot Concepts - A Powerful Technique in Well Log
Analysis. JPT March, pp 535-549.

Fertl W.H, 1983 Gamma Ray Spectral Logging: A New Evaluation Frontier, Pt 6, Clay
Analysis in Shaly Sands. World Oil Vol 197 No 5 pp 99-112.

Hamouz M.A. & Mueller S.L. 1984 Some New Ideas for Well Log Pore-Pressure
Prediction. SPE 13204.

Herring E.A. 1973 Estimating Abnormal Pressures from Log Data in the North Sea. 2nd
Annual SPE of AIME Europe Mtg Reprint No. SPE 4301, 8 pp.

Hottman C.E. 1965 Estimation of Formation Pressures from Log-Derived Shale


Properties. JPT Vol 17, June, pp 717-722.

Lane R.A. & McPherson L.A. 1976 A Review of Geopressure Evaluation from Well Logs -
Louisiana Gulf Coast. JPT Vol 28 Sept, pp 963-971.

McKee R.E. & Pilkington P.E. 1974 Pressure Prediction and Detection Conclusion: If in
doubt, log to confirm overpressures. Oil & Gas Journal, Vol 72 No 51 pp 29-31.

Overton H.L. and Timko D.J. 1969 The Salinity Principle, A Tectonic Stress Indicator in
Marine Sands. The Log Analyst, Vol 3, May-June, pp 34-43.

Pennebaker E.S. 1968 Seismic Data Indicate Depth, Magnitude of Abnormal Pressures.
World Oil No. 166, pp 73-78.

Wireline Logs 5-17 Ver. 1.0


Halliburton Overpressure

Reynolds E.B., Timko D.J. & Zanier A.M. 1973 Potential Hazards of Acoustic-Log Shale
Pressure Plots. JPT Sept, pp 1039-1044.

Waters S. & Moore N. 1978 Pore Pressure Predictions from High Resolution Seismic Data.
10th Annual SPE of AIME Offshore Technical Conference. Reprint
No. OTC 3220, pp 1443-1454.

Zoeller W.A. 1983 Pore Pressure Detection from the MWD (Measurement While Drilling)
Gamma Ray. 58th Annual SPE of AIME Technical Conference.
Reprint No. SPE 12166, 16 pp.

Wireline Logs 5-18 Ver. 1.0


Halliburton Overpressure

SECTION 6
OVERPRESSURE DETECTION - SEQUENCE FOR
WELLSITE ANALYSIS
6.1 BEFORE DRILLING
Overpressure detection must begin prior to drilling. It is imperative that some information is
calculated before drilling commences if good understanding of the pressure regime is to be
gained. Input all data into the relevant files in the wellsite computer, and process as if the data
has been obtained from the current well.

1 Obtain as much information from any adjacent wells as possible. This information
should include wireline logs, pressure logs, and mud logs.

2 From wireline logs calculate average densities. The density log (FDC or equivalent)
should be used if available. If not, the sonic log may be used.

3 Plot an overburden gradient curve for the area (use local air gap, sea depth, and
water density). Synthesise the data input from all available wells, using the prognosis
of the current well to supply information on formation type depths and thicknesses.

4 Plot the following

a) Shale sonic transit time


b) Shale resistivity (ILD or LL deep)
c) Shale density
d) dc exponent.

All these plots must be drawn on the same depth scale for easy correlation.

5 Use any direct formation pressure readings (ie RFT, DST, well kicks) and mark on the
above plots. Using this information, back-calculate the Normal Compaction Trend
(NCT).

6 If there are no direct pressure readings available, estimate the NCT and construct an
Eaton overlay grid. Draw an estimated pore pressure profile alongside the plots.
Compare these pressure plots for agreement or disagreement.

7 Consider the possibility of the prognosed well being similar or dissimilar in its
pressure profile compared to offset wells. For example, do prognosed depths agree?

8 Note whether the NCTs are obvious, shifted by unconformities, whether the regime is
undercompacted, has a physical or chemical seal, etc. Keep this information and plots

Sequence for Well Analysis 6-1 Ver. 1.0


Halliburton Overpressure

near at hand so that comparisons with the current well can be made as drilling
progresses.

9 Obtain, if possible, seismic interval velocities. Plot as for sonic information, and draw
on NCT. Estimate tops and magnitude of pressures.

This information should provide sufficient local knowledge to formulate a good estimate of
formation pressures.

6.2 PRESSURE EVALUATION DURING DRILLING


During drilling, the overpressure log must be constantly updated. Information recorded
should include

a) dc exponent
b) Cuttings Density
c) Shale Factor
d) Mud Temperature Out, including an end-to-end trend plot
e) Differential Temperature
f) Gas. Ensure that the actual gas chart recorder is watched, not just the logging
worksheet, as connection gases may be critical
g) Drilling information, including torque, cuttings, overpull and drag, and also
the mud rheology.

Once this information is plotted, draw on NCTs and construct overlays. Believe the
indications, even if they do not agree with other wells in the area. Do not hesitate to redefine
estimates (where necessary) as more information becomes available.

Kicks will always accurately define NCT

The above should be followed at all stages of drilling until Total Depth (TD) is reached.
Always remember to re-evaluate results, if necessary, as more information becomes available.

6.3 AFTER DRILLING


At TD the following parameters should be noted

1 RFT results
2 Formation pressures from DSTs

If estimated pressures were incorrect, attempt to discover why. Any new methods, pitfalls,
techniques, etc introduced by the client, drilling contractor, or any other service companies
should be passed on to the Halliburton office.

Sequence for Well Analysis 6-2 Ver. 1.0


Halliburton Overpressure

SECTION 7
FRACTURE PRESSURE
7.1 FRACTURE GRADIENTS
Knowledge of the magnitude of formation fracture gradients is vital, especially when drilling
into an abnormally pressured zone. Formation fracture gradient determines the maximum
allowable mud weight that can be used (after incorporating an operational safety factor).

7.1.1 Mechanism of Formation Fracturing


A formation can be made to fracture by the application of fluid pressure to overcome the least
line of resistance within the rock structure. Normally fractures will be propagated in a
direction perpendicular to the least principal stress. Which of these three stresses is the least
can be predicted by the fault activity in the area.

Figure 7-1: Distribution of Stress Planes in Unit of Rock Formation

Normal faulting indicates that the least principal stress is horizontal and probably about
equal to the minimum horizontal stress required to avoid rock failure (figure 7-2).

Figure 7-2: Normal Fault

Fracture Pressure 7-1 Ver. 1.0


Halliburton Overpressure

Thrust (reverse) faulting indicates the least principal stress to be of a vertical nature. It must
be at least equal to the weight of the overburden, or about 1 psi/ft (figure 7.3).

Figure 7-3: Reverse Fault

Transcurrent faulting indicates that the least principal stress is horizontal but can be larger
than the minimum required to avoid failure. It cannot be larger than the vertical stress
(figure 7-4).

Figure 7-4: Transcurrent Fault

To initiate the fracture fluid pressure must be transmitted to the formation, and to propagate
the fracture this pressure must be maintained at a level greater than the least principal stress.
If mud is used, then the fluid pressure has to overcome the filter cake deposited on the
borehole wall, plus the pressure at which the formation fractures. It is therefore preferable to
use a clear, low viscosity fluid when making a determination of the formation breakdown
gradient.

In practice, formation breakdown gradients are determined by performing a Leak Off Test
(LOT). This is normally done after drilling through a casing shoe in order to determine the
maximum allowable mud gradient for the next section of open hole. The usual procedure is
outlined below.

a) Drill through the casing shoe and into new formation to a depth of 15 to 20 feet. Close
the blowout preventers.

b) Raise the surface pressure in increments until bleed off is indicated. Alternatively,
pumping could be maintained at a slow rate until bleed off.

Fracture Pressure 7-2 Ver. 1.0


Halliburton Overpressure

Figure 7-5: Typical Leak Off Test

During fracture testing it is generally not the coherent rock matrix that is sheared, rather the
existing natural discontinuities (joints, parting, etc) which are normally held together by the
compressive stresses in the formation. When the hydraulic stress in the borehole equals the
formation compressive stress, it effectively reduces the latter to zero across the crack. If the
borehole pressure is then raised slightly, fluid will enter the crack.

7.2 LEAK OFF TESTS AND FRACTURE PRESSURES


Terzaghi and Peck (1948) states that
S=+P

where S = Overburden Stress


= Matrix Stress
P = Pore Pressure

As previously stated, subsurface stress is characterised by three unequal, mutually


perpendicular principal stresses.

h) is of greatest
The greatest stress is vertical (v) although the least principal stress (
interest as this indicates formation strength along the weakest plane.

7.2.1 Hubbert and Willis (1957)


Hubbert and Willis conducted laboratory experiments on core samples and derived the
empirical relationships below.

Fracture Pressure 7-3 Ver. 1.0


Halliburton Overpressure

To extend a fracture

1
h = v
3

To induce a fracture

1
h = v
2

Hence
or
Therefore, to extend 1 1
1 v < h < 1 v
while to induce (S3 - P) < h < 2 (S - P)
Therefore to fracture a formation3 one 1 2both the pore pressure and the
FPmust
= Povercome
+ (S - P)
horizontal matrix stress of the rock. FP = P + 31 (S - P)
2
This value (1/3 to 1/2) is only an estimate. The actual value will depend upon lithology and
compaction of the sediments. To take account of this, the Poisson's Ratio was introduced.

7.2.2 Poisson's Ratio


From Young's Law

where n=1 h n
= ( )
= Poisson's Ratio v 1 -

Hubbert and Willis estimated Poisson's Ratio as between 0.33 and 0.25, ie

0.33 describes
Poisson's Ratio is a property which 1 the behaviour
0.25 of rock 1 stresses in the least
= while =
principal stress direction when 0.33 is2applied in
1 -pressure 1 the
- 0.25
direction3 of the principal stress. For
elastic rocks, the ratio of the stresses is represented by

Laboratory tests on unconsolidated rockshave


L
= shown that
P 1 -
L
Field tests have shown to have values=between 1
ie 0.25 and 0.5, above which the rock becomes
= 0.25
3
plastic, ie the stresses are equal in allP directions. Poisson's Ratio varies with both rock type
and the degree of compaction, and will never exceed 0.5. Every rock will have its own
characteristic Poisson's Ratio.

Note should be made of unconsolidated clay formations, often found at shallow depths in
North Sea wells, which may exhibit abnormally high fracture pressures. Wet clays will
behave plastically and the Poisson's Ratio will approach 0.5. Due to the pore water and
adsorbed water surrounding the individual clay platelets, these platelets will not be in contact
with each other, and so these clay types will have very low shear strength. Thus the pore

Fracture Pressure 7-4 Ver. 1.0


Halliburton Overpressure

water would be supporting the weight of the overlying sediments and the pore pressure should
almost be equal to the overburden pressure. As a result, hydraulic pressures greater than the
overburden pressures would be necessary to fracture this formation. Under these
circumstances, a horizontal fracture would form lifting the overburden.

With increased depth, compaction will squeeze out some of the pore water, bringing the clay
platelets into contact. They are then able to support a superimposed horizontal stress, and
Poisson's Ratio will become similar to that of a more compact clay.

To build up a complete fracture gradient picture for any well, the fracture pressure should be
calculated every time a lithology or pore pressure change is noted. This is particularly
important when drilling into sands or zones of decreasing pore pressure gradient, which may
be less competent than the formation at the shoe.

In the case of changing lithology, evaluation is best carried out using experimentally derived
Poisson's Ratio values for different lithologies, such as those of Weurker (1963) shown in
table 7-1.

7.2.3 Matthews and Kelly Method


For compacted shales, it has been shown that the overburden stress (S) is supported by the
formation fluid pressure (P) and the grain-to-grain bearing strength () of the clay particles.
S=+P

In the above equation represents the matrix stress, and it is evident that it's value will
depend solely upon the degree of rock compaction. Hubbert and Rubey state that "a useful
measure of the degree of compaction of a clay is its porosity". This is defined as the ratio of
pore volume to the total volume. From this it may be inferred that, for any given clay, for
each value of , there is some minimum value of matrix stress which the clay can support
without further compaction. The porosity at a given burial depth (D) is dependent on the
fluid pressure (P). Thus if the fluid pressure is abnormally high (greater than hydrostatic) the
porosity will also be abnormally high, and so the matrix stress would be abnormally low. The
converse is assumed to be true if the formation fluid pressure is abnormally low.

It has been assumed that whenever a formation is fractured horizontally the fracture pressure
is about equal to the overburden stress. However, for the case of a vertical fracture, the stress
required may be much less then the overburden load at the depth where the fracture is created.
Nevertheless, in both the horizontal and vertical fracture cases, it may be safely assumed that
the fracture will not begin until the applied pressure is at least equal to the formation fluid
pressure. If this is the case, then any additional pressure required may be related to the matrix
stress and hence will only vary with the degree of compaction. The relationship below has
been developed for calculating the fracture gradient of sedimentary formations.

Fracture Pressure 7-5 Ver. 1.0


Halliburton Overpressure

P
F = +
D D

where P = the formation fluid pressure (psi) at depth of interest D


= the matrix stress (psi)
(chi) = the matrix stress coefficient for the depth at which the value
of would be the normal matrix stress
F = the fracture gradient in psi/ft

Matthews and Kelly (1967) assumed 1.0 psi/ft for overburden, but replaced the Poisson's
).
Ratio by a matrix stress coefficient (

was determined from empirical data, ie leak off tests.

FP P P
= + (1 - )
D D D

This showed that

a) was depth dependent


b) was geographical area dependent

7.2.4 Eaton (1969)


Eaton established variable overburden gradients from bulk densities, and calculated a
Poisson's Ratio from empirical data.

FP S P P
= ( - )( )+
D D D 1 - D

where S/D from density logs


P/D = actual pore fluid pressure
from empirical data, or by Anderson et al method for sands

7.2.5 Anderson et al (1973)


Anderson expands Biot's (1955) stress/strain relationship

FP 2 S 1 - 3 P
= [ ] + [a ( ) ]
D 1 - D 1 - D

Fracture Pressure 7-6 Ver. 1.0


Halliburton Overpressure

where

Cr
a = 1 -
Cb

Cr = intrinsic compressibility of solid rock material


Cb = compressibility of solid rock skeleton

Cr/Cb = function of shale estimated in cuttings

Fracture Pressure 7-7 Ver. 1.0


Halliburton Overpressure

If cuttings contain 100% shale then a = 0, ie no porosity, hence pore pressure does not have
an effect. The equation becomes

FP 2 S
=
D 1 - D

7.2.6 Daines (1982)


Daines states that there are two unequal horizontal stresses which must be overcome before
fracturing occurs.

1 That caused by the weight of overlying sediments (h)

2 A superposed tectonic stress (T)

Proof that this tectonic stress exists is evidenced by folding, faulting, etc, which rely on
unequal stress states for their occurrence and maintenance.

The minimum horizontal stress occurs when the tectonic stress (T) is zero, ie


H = V ( ) where V = (S - P)
1 -

The superposed horizontal tectonic stress will increase uniformly with depth such that the
ratio T:v remains constant.

The value of T is calculated from the first LOT by subtracting the Eaton calculated fracture
pressure from the LOT result, ie


T = LOT - [ V ( ) + P]
1 -

where V = S - P

The ratio of T:V is then calculated so that for subsequent points in the borehole the value of
T can be calculated from

= Vi ( Tl )
l

T
i

where i denotes depth of interest


l denotes LOT depth

Fracture Pressure 7-8 Ver. 1.0


Halliburton Overpressure

As a result, the fracture pressure can be calculated for any depth if the overburden gradient,
pore pressure, and Poisson's Ratio for the lithology at that depth are known


FP = T + [ V ( ) + P]
1 -

To simplify this calculation Daines paper provides a list of typical Poisson's Ratios for certain
lithologies reproduced here in table 7-1 (from Weurker 1963).

Obviously if this data is to be accurate then strict lithological control is necessary. Problems
arise when two types of lithology are intermixed eg sandy clay and shaly sand. In these cases
the mineral forming the matrix must be determined and the Poisson's Ratio for that used.

Daines quotes a figure of 95% accuracy for use of his equations.

Fracture Pressure 7-9 Ver. 1.0


Halliburton Overpressure

Formation Type Poisson's ratio


clay, very wet 0.50
clay 0.17
conglomerate 0.20
dolomite 0.21
greywacke:
coarse 0.07
fine 0.23
medium 0.24
limestone:
fine, micritic 0.28
medium, calcarenitic 0.31
porous 0.20
stylolitic 0.27
fossiliferous 0.09
bedded fossils 0.17
shaly 0.17
sandstone:
coarse 0.05
coarse, cemented 0.10
fine 0.03
very fine 0.04
medium 0.06
poorly sorted, shaly 0.24
fossiliferous 0.01
shale:
calacareous (<50% CaCO3) 0.14
dolomitic 0.28
siliceous 0.12
silty (<70% silt) 0.17
sandy (<70% sand) 0.12
kerogenaceous 0.25
siltstone 0.08
slate 0.13
tuff (glass) 0.34

Table 7-1: Suggested Values of Poisson's Ratio

Fracture Pressure 7-10 Ver. 1.0


Halliburton Overpressure

(From Weurker R.G. "Annotated Tables of Strength & Elastic Properties of rocks," Drilling
Reprint Series SPE, Dallas (1963))

7.2.7 Poisson's Ratio Calculations


This method was described by Tony Taylor (1978) and is based on an equation formulated by
Tixier, Loveless and Anderson in 1973. The Poisson's Ratios calculated by this method can
be applied to sands, particularly those with significant clay content.

Since the minimum value of available is 0.27, this method cannot be used to define
Poisson's Ration in finer grained clastics. Also, since carbonates tend to have low values of
gamma ray the Poisson's ratio would appear much higher than would be appropriate for such
competent rocks.
= 0.125 q + 0.27

where = Poisson's Ratio


0.125 = constant
q = Shaliness Index
0.27 = theoretical minimum for Poisson's Ratio in a clean sand

To calculate shaliness index (q) use the gamma ray log. Average the readings over 50 ft
intervals and insert into the formula

GRlog - GR min
q =
GR max - GR min

This value is substituted into the equation for above, and in turn Poisson's ratio () may be
substituted into the desired fracture gradient equation discussed above.

Fracture Pressure 7-11 Ver. 1.0


Halliburton Overpressure

REFERENCES
Anderson R.A. Ingram D.S. & Zanier A.M. 1972 Fracture Pressure Gradients
Determination from Well Logs. 47th Annual SPE of AIME Fall Mtg,
Reprint No. SPE 4135, 15 pp.

Biot M.A. 1955 Theory of Elasticity and Consolidation for 2 Porous Anisotropic Solids
J. Appl. Phys. Vol 26 No 2 pp 115-135.

Cesaroni R. Giacca D. Schenato A. & Thierree B. 1981 Estimation of Overburden and


Fracture Gradients in Clastics from Drilling Parameters On-Site Processing. Pet. Eng. Intl.
June pp 60-86.

Christman S.A. 1973 Offshore Fracture Gradients and Casing Setting Depths. JPT August,
pp 910-914.

Daget P. & Parigot P. 1979 Using Log Data to Predict Leak-Off Test Pressures. World Oil
Vol 188 No 2 pp 48-52.

Daines S.R. 1982 Prediction of Fracture Pressures for Wildcat Wells. JPT Vol 34 No 4
pp 863-872.

Eaton B.A. 1969 Fracture Gradient Prediction and it's Application. Oilfield Operations.
JPT October, pp 1353-1360.

Fertl W.H. 1976 Predicting Fracture Pressure Gradients for More Efficient Drilling.
Petrol Eng., Vol 48 No 14 pp 56-71.

Hubbert M.K. & Rubey W.W. 1959 Role of Fluid Pressure in Mechanics of Overthrust
Faulting. Geol Soc Amer Bull (Feb 1959) Vol 70 pp 115-206.

Hubbert M.K. & Willis D.G. 1957 Mechanics of Hydraulic Fracturing. Trans AIME Vol
210, pp 153-166.

Matthews W.R. & Kelly J. 1967 How to Predict Formation Pressure and Fracture
Gradient. Oil & Gas Journal Feb, pp 92-106.

Nolte K.G. & Smith M.B. 1981 Interpretation of Fracturing Pressures. SPE of AIME
Rocky Mt Reg Mtg (Casper, Wyo) Reprint No. SPE 8297, 8pp.

Prats M. 1981 Effect of Burial History on the Subsurface Horizontal Stresses of Formations
Having Different Material Properties. SPEJ Vol 21 No 6 pp 658-662.

Taylor A.H.1978 NEC Gas - Internal Memo.

Fracture Pressure 7-12 Ver. 1.0


Halliburton Overpressure

Tixier M.P., Loveless G.W. & Anderson R.A. 1973 Estimation of Formation Strength from
Mechanical Properties Log. 48th Annual SPE of AIME Fall Mtg,
Preprint No. SPE 4532, 14 pp.

Weurker R.G. 1963 Annotated Tables of Strength and Properties of Rocks. Drilling SPE
Petroleum Trans Reprint Series No. 6.

Fracture Pressure 7-13 Ver. 1.0


Halliburton Overpressure

Fracture Pressure 7-14 Ver. 1.0

You might also like