Sie sind auf Seite 1von 28

Chapter

1
Slurry Design
Technology for cementing deep wells has advanced greatly in recent years.
Operational conditions once considered difficult or impossible are now dealt with
frequently. Many wells are now deeper than 15,000 ft, and deep wells with
bottomhole temperatures above 230F are always considered critical. The basic
cementing procedures used for high-pressure/high-temperature (HPHT) wells
are similar to the cementing procedures used in shallower and cooler wells, but
HPHT wells require even greater emphasis on cement slurry design.

Each well has certain basic characteristics that dictate the required cement slurry
properties and performance. Careful and thorough review of these well charac-
teristics is essential for designing an effective slurry. The overall goal of design-
ing a cement slurry for a specific well application is to select an economical
cement mixture that can be placed under existing well conditions. This slurry
should develop and retain the properties necessary to isolate zones as well as
support and protect the casing or liner.

Designers should combine examinations of individual cementing variables to


develop a total cement job design. For example, to properly design cement
slurries for hostile gas wells, designers must thoroughly understand the mecha-
nism that causes loss of hydrostatic pressure in a cement column. In addition,
they must (1) carefully design for fluid-loss control, cement stability, and setting
behavior, and (2) examine drilling fluid conditioning and spacers.

Unless proper zonal isolation can be achieved, it will be impossible to indepen-


dently produce from the different reservoirs the well penetrates. In the case of
faulty zonal isolation, it will be impossible to perform chemical treatments in the
necessary intervals. Faulty zonal isolation also will result in fluid migration in the
annulus. The remedial cementing necessary to correct uncontrolled fluid flow
behind casing is time-consuming and expensive. Remedial cementing also
weakens the integrity of the casing.

April 1996 Slurry Design 1-1


This section examines the factors involved in designing a slurry for HPHT wells
by covering the following topics:

Wellbore Temperatures
Retardation
Density
Filtration Control
Strength Stability
Viscosity/Suspension
Gas Migration Theory and Control
Cement Job Simulation
Waiting on Cement Time

Wellbore Temperatures
One of the most important factors controlling the chemical reaction and perfor-
mance results of a cementing composition is the wellbore temperature. In oilwell
cementing, the cement slurry is subjected to progressively increasing tempera-
tures from the time it is mixed on the surface and pumped into the well until the
time the cement cures and the formations adjacent to the wellbore return to their
ultimate static temperature. Circulating and static temperature both affect
cement design. Circulating temperature refers to the temperature the slurry
encounters as it is being pumped into the well. Static temperature refers to the
formation heat the wellbore fluids will be subjected to after circulation is stopped
for a set period of time. Although static temperatures affect the curing properties
of the cement, circulation temperature has an even greater influence.

Bottomhole Circulating Temperature


Bottomhole circulating temperature (BHCT) is the temperature that influences
the thickening time or pumpability of the cement slurry. The BHCT is normally
calculated from a set of temperature schedules published in API Specification
10. Recently, several different temperature subs have been used in many wells in
the United States. These wells have provided detailed information that was not
available when the first circulating temperature relationship was developed. As a
result, API has used this new set of data to develop a better overall correlation
for calculating BHCT. Several companies are already using this alternate calcu-
lation method, and it soon will be listed in API RP 10B. The new API correla-
tions already are incorporated in CJOBSIMTM (version 2.2.1) and CEMLAB
(version 1.0.3).

1-2 High-Pressure/High-Temperature Cementing April 1996


Since accurately estimating BHCT is essential, designers should not rely on
the API schedules alone when cementing deep wells. Temperatures should be
verified by some form of actual downhole measurement, preferably during
the circulation phase. Halliburton has developed the BHCT II recorder, a
temperature gauge that can reliably measure and record BHCT (Figure 1.1).

Bottomhole Static Temperature


Knowing the bottomhole static temperature (BHST) is important for design-
ing and assessing long-term stability or rate of compressive strength develop-
ment for a cement slurry. Determining BHST is especially important in deep-
well cementingwhere the temperature differential between the top and
bottom of the cement can be high. For example, cement slurries that are
designed for safe placement times may be overretarded at top-of-cement
(TOC) temperatures, resulting in poor compressive strength development.
Generally, if the static temperature at the top of the cement column exceeds
the BHCT, overretardation is not expected.

Precise temperature readings are essential for cementing purposes. An error


as small as 5 to 10F can significantly affect results even though slurry
materials are chosen to minimize the effect of slurry sensitivity, however.
Properties of some cement systems, and the conditions in which they are
applied, dictate the need for the designer to know wellbore temperature.
Generally, cement sensitivity increases as BHCT increases. As a result, all
laboratory tests performed to improve slurry properties should be run using
samples of the same batch of cement, mixing water, and chemical cementing
additives that will be used during the job. Downhole conditions must be
duplicated as closely as possible. The best method of estimating bottomhole
temperatures in horizontal or deviated wells is to simulate temperatures with
the Enertech temperature simulator.
Figure 1.1BHCT II
recorder.
Retardation
Thickening Time
The high temperatures in deep wells can cause cement to set prematurely
unless the slurry has been properly retarded. In addition, the pressure imposed
on a cement slurry by the hydrostatic load of well fluids tends to accelerate
hydration, thereby reducing the pumping capability of the cement.

The cement design must contain just enough retarder to delay the slurry from
setting long enough to place the slurry downhole. To prevent under- or
overretarding the slurry, an accurate BHCT must be obtained for laboratory
testing. Specific thickening time recommendations depend largely on the type of
job, the well conditions, and the volume of cement being pumped. Temperature
and pressure both influence cement setting; however, pump rate, casing size, and
well depth determine the placement time.

April 1996 Slurry Design 1-3


Generally, the cement slurry thickening time for a HPHT well should be ap-
proximately double the placement time. The slurry should be designed so that
the viscosity of the cement stays reasonably low during three-fourths of the
thickening time test. If slurry viscosity increases prematurely during placement,
the resulting high friction pressures could possibly break down the formation
(Figure 1.2). Slurry thickening time should not be sensitive to small changes in
retarder concentration. A moderate increase in retarder concentration should only
have a moderate effect on the slurrys thickening time.

Humps
Undesirable
Viscosity

Low Viscosity
Fast Setting

Time

Figure 1.2Thickening time behavior in deep wells.

Slurry Design
For most deep-well casing jobs, the cement slurry is designed to have a 4- to 5-
hour pumping time. This length of time normally allows an adequate safety
factor, since few cementing jobs, even large ones, require more than 2 hours to
place the slurry. Setting and cementing 6,000- to 8,000-ft liners in many deep-
well drilling areas is now common, and any cementing composition that covers
this distance must be retarded enough to be placed at the higher bottomhole
temperatures while still developing reasonable strength at the lower temperatures
across the liner lap (Figure 1.3). A temperature difference exceeding 100F from
top to bottom is not uncommon. Excessively retarding cement slurries to achieve
greater thickening times also increases the possibility of gas cutting. In addition,
it may impair an otherwise successful cement job around the top of the liner. On
deep liner jobs having fairly high downhole temperatures, the 4- to 5-hour
pumping time is still usually adequate.

1-4 High-Pressure/High-Temperature Cementing April 1996


When whipstock plugs are spotted, their thickening times should
be as short as safely possible. In squeeze cementing, the fluid
time requirements for different techniques may vary. Shutdowns
to simulate a hesitation-type squeeze during thickening time
tests have shown significant reductions in the pumping capabil-
ity of a cement slurry.
Top of Liner Static
Cement Retarders
Halliburtons HR-12 and HR-15 cement retarders are lignosul-
fonate additives that are commonly used in deep-well cementing. Temperature
Differential
SCR-100 cement retarder is a synthetic polymer, therefore its
composition is more consistent than natural by-products such as
lignosulfonates. Also, its thickening time response is less
sensitive to cement variations as compared to other classic
Bottomhole Circulating
retarders. SCR-100 is especially effective when good short-
term compressive strength development is necessary at the top
of a long liner. HR-25 is used to increase the retarding proper- Figure 1.3Strength development at the
ties of SCR-100 above 250F. top of long liners.

Salt Cement
Invert-emulsion drilling fluids are used in many deep, hot wells. A high concen-
tration of salt is usually contained within the internal aqueous phase of these
drilling fluids. This salt can pose compatibility problems with some cement
slurries. As Table 1.1 shows, contamination tests conducted with freshwater
designs have considerably shorter thickening times and extremely low compres-
sive strength development after 24 hours. In contrast, saturated salt cements tend
to reduce the contamination effect of these oil-based drilling fluids. In several
cases, cement designs containing 18% salt have been found to be just as effec-
tive as saturated salt designs. As a result, it is sometimes necessary to complete
these wells with a cement slurry containing salt concentrations up to saturation.
Even if salt cement designs are not used in these situations, the additives used
should be relatively salt-tolerant.

Table 1.1Contamination Effect on Invert Emulsion Muds


Thickening
24-hour Compressive Strength
18 lb/gal Drilling Time
Cement Designs Fluid (hr:min) psi % of base slurry
Saturated Salt None 6:50 6,790 --
Saturated Salt NaCl 6:18 2,010 30
Saturated Salt CaCI2 3:44 1,790 26
Fresh Water None 4:00 1,900 --
Fresh Water NaCl 2:22 80 4
Fresh Water CaCI2 1:45 130 7

April 1996 Slurry Design 1-5


Density
To cement across high-pressure gas zones, a high-density drilling fluid system
must be used to control gas until the well can be cased off and cemented. These
high-pressure gas sections require drilling fluid weights up to 20 lb/gal. With the
exception of a squeeze design, cement slurry designs should always have a
greater density than the drilling fluid system, unless lost circulation is a concern.
Even then, cement density should never be lower than drilling fluid density.
Higher densities will increase the overbalance pressure adjacent to the gas zone,
which can help prevent gas migration. Ideally, cement density should be at least
1 lbm/gal heavier than drilling fluid density.

For heavyweight cement designs, slurries will commonly use heavyweight


materials with low water requirements such as SSA-2 coarse silica and HI-
DENSE No. 3 weight additive. While HI-DENSE is chemically inert, some
form of silica (SSA-2) is required in high-temperature cementing to prevent
strength retrogression. CFR-3 dispersant also can be used in heavyweight
cement designs to improve surface mixing or to lower the water-to-cement ratio,
which increases slurry density.

For proper cement performance, a slurry must be maintained at its correct design
density. Since heavyweight slurries are often fairly viscous when mixed at
surface conditions, they tend to entrain air, especially when mixed with salt or
seawater. This characteristic can make it difficult to achieve the necessary slurry
density unless proper antifoam agents are used. Using a pressurized drilling fluid
scale (Figure 1.4) can help ensure that the cement slurry pumped downhole has
the proper density.

The following properties have the


greatest effect on density.

slurry volume
flow characteristics
thickening time
free water
suspension
fluid-loss control

Any one of these characteristics can


affect the results of a primary cement
job.

Figure 1.4Pressurized drilling fluid scale.

1-6 High-Pressure/High-Temperature Cementing April 1996


Filtration Control
In any deep-well cementing job, the cement slurry should contain fluid-loss
additives that perform well at high temperatures. Fluid-loss additives help
maintain designed slurry properties by preventing cement dehydration. Good
fluid-loss control also offers important protection during primary cementing,
particularly where casing clearance is small (Figure 1.5) or where highly perme-
able sections are encountered (Figure 1.6 on Page 1-8). Filtration control also
can prevent the following potential cementing problems.

high friction pressures or premature shutdown


damage to water-sensitive formations
annular gas migration

Typical Liner-Hole Combinations


Casing (in) 9 5/8 7 5/8 7 5/8 7 7 5 1/2
Hole (in) 8 5/8 6 5/8 6 5/8 6 1/8 6 1/8 4 3/4
Liner (in) 7 5 1/2 5 5 4 1/2 3 1/2

Annulus Area - (Sq in) 19.9 10.7 14.9 9.9 13.6 8.1
Cement Sheath Thickness 13/16 9/16 13/16 9/16 13/16 5/8
Illustrations

Figure 1.5Casing clearance and resulting annulus area.

Guidelines for fluid-loss control vary with area and experience. For cementing
casing, a fluid loss of 150 to 300 cc/30 min is typically used. For liner and
squeeze systems, a fluid loss of 50 to 100 cc/30 min is common. In areas where
potential for fluid migration exists, a fluid-loss value of less than 50 cc/30 min is
normal practice. These values are recommendations only; experience in specific
areas will dictate proper filtration control. Good fluid-loss control also is neces-
sary for successful squeeze cementing. In this application, uncontrolled fluid
loss can result in rapid cement dehydration, which can bridge off the wellbore
before the entire zone of interest is squeezed (Figure 1.7 on Page 1-8). If fluid
loss is controlled, cement can contact the entire interval, allowing small nodes of
dehydrated cement to build up across permeable areas of interest before squeeze
pressure develops (Figure 1.8 on Page 1-9).

April 1996 Slurry Design 1-7


Primary Cementing

Permeable
Zone

Thin Heavy Cement


Impermeable Filter Cake
Cement
Filter Cake

Low Uncontrolled
Fluid Loss Fluid Loss

Figure 1.6Highly permeable sections require good fluid-loss


control.

Cement
Dehydrated
Cement
Mud Filter
Cake

Mud

Figure 1.7Uncontrolled fluid loss can result in rapid cement


dehydration.

1-8 High-Pressure/High-Temperature Cementing April 1996


Squeeze Cementing

1,000cc/
Cement 30 min.
Filter Cake
300cc/
30 min.

150cc/
30 min.
25cc/
30 min.

Filter Cake Buildup Inside Casing

Figure 1.8Cement node buildup.

In high-temperature cementing jobs, HALAD-413 fluid-loss additive,


HALAD-36lA cement additive, FDP-C471, HALAD-100, and HALAD 600
LE offer excellent fluid-loss properties. LATEX 2000 additive also can be used
to provide excellent downhole fluid-loss control.

Strength Stability
Temperatures in the range of 400 to 500F for deep oil and gas wells are not
uncommon. Geothermal wells with bottomhole temperatures ranging from 400
to 750F producing flashing brine are frequently encountered. Cement slurries
used in these operations should remain competent at high temperatures and
pressures for extended periods. Set cement exposed to temperatures above
230F gradually increases in permeability and decreases in strength. Strength
retrogression may eventually provide inadequate pipe support, and increased
permeability may result in the loss of zonal isolation (Figure 1.9 on Page 1-10).

When Portland cement is mixed with water, tricalcium silicate (C3S) and
dicalcium silicate (C2S) undergo hydration to form calcium silicate hydrate
(C-S-H) gel and hydrated lime Ca(OH)2. At temperatures above 230F,
C-S-H gel converts to -dicalcium silicate hydrate (-C2SH). Conversion to the
-C2SH phase results in a loss of compressive strength and an increase in
permeability. Conversion of C-S-H gel to -C2SH at 230F and above can be
prevented by the addition of crystalline silica (fine sand: SSA-1 or coarse sand:
SSA-2). Concentration of crystalline silica values range from 35 to 70% depend-
ing on the bottomhole static temperature and the water content of the slurry. The

April 1996 Slurry Design 1-9


minimum concentration (35%) should be achieved; however, the exact concen-
tration used depends on the required slurry properties. In most cases, 35 to 40%
crystalline silica will meet design criteria. Recent work has shown that silica
flour (SSA-1) provides maximum benefit for temperatures ranging from 230 to
440F. For conditions where density is critical, coarse grade sand (SSA-2) is
preferred for densified slurry. The reaction between SSA-2 and cement takes
place slower than between SSA-1 and cement.

4000
Coarse Silica
Compressive Strength (psi)

3000 Fine Silica

2000

1000
No Silica

0
0 8 16 24 32
Time (hours)

Figure 1.9High-temperature compressive strength response.

SSA-1 or SSA-2 increases the molar ratio of CaO/SiO2 to 1 or less, resulting in


the formation of crystalline tobermorite. Formation of the tobermorite phase
prevents strength retrogression and permeability increase. At temperatures above
300F, tobermorite primarily converts to xonotlite. At 480F, truscottite begins to
appear. Above 750F, xonotlite and truscottite reach the limit of their stability. At
higher temperatures, the crystalline phases xonotlite and truscottite dehydrate,
resulting in the breakdown of the set cement. To achieve maximum stability at
high temperatures, a CaO/SiO2 ratio of 1 or less should be maintained.

Viscosity/Suspension
In primary cementing, a cement slurry should have a reasonable viscosity during
mixing operations and should maintain good suspension properties under
downhole conditions. Suspension properties are important because solids
segregation causes fluid to migrate upward through the slurry after it is placed in
the wellbore, creating an uncemented area in the annulus. Free fluid is particu-
larly harmful in deviated or horizontal wells where segregation/separation is
more likely. This free fluid can collect along the high side of the annulus and
form a channel that contributes to zonal communication or gas migration. Even

1-10 High-Pressure/High-Temperature Cementing April 1996


in vertical wells, pockets of free fluid located near a corrosive water zone can
eventually result in casing leaks. Therefore, it is important to design slurries that
are stable under downhole conditions.

In most critical applications, a slurry should be designed with no measurable


free water. Settling tests should be conducted after the cement has set; these tests
were developed by British Petroleum (BP) and have been approved as a recom-
mended practice by API. Density variations of less than 0.5 lb/gal along an 8-in.
column are acceptable test results.
Viscosity

Temperature

Figure 1.10Thermal thinning effect.

It is more difficult to formulate a cement design with good suspension properties


for high-temperature conditions. The downhole thermal thinning effect makes it
difficult to design a slurry with a reasonable viscosity during mixing that will
still have enough suspension properties downhole to prevent solids segregation
(Figure 1.10).

Another factor involves complex slurry designs that are often used in high-
temperature applications such as liner jobs. These slurries are susceptible to
solids segregation because they commonly use dispersing additives such as
CFR-3 dispersant or lignosulfonate-based retarders. Since these additives tend to
reduce the low shear rheology, yield point, and static gel strength development
of the slurry, they increase the potential for solids segregation. This effect can be
offset by using viscosifying fluid-loss additives such as cellulose derivatives.

Since several of the fluid-loss additives used in high-temperature cementing


applications are nonviscosifying (i.e., HALAD-361A and HALAD-413), designs
incorporating these additives often must be supplemented with viscosifying
agents to maintain good solids suspension downhole. In contrast, LATEX-2000

April 1996 Slurry Design 1-11


cement not only imparts excellent fluid-loss control but also provides good
suspension properties to the slurry without increasing its viscosity.

High-temperature environments often require the use of a heavyweight additive


such as HI-DENSE No. 3 to control formation pressures. While this additive
contains a fairly high concentration of coarse particles and has a high specific
gravity, excellent suspension properties are still necessary to prevent settling.
Another heavyweight additive with a similar specific gravity is HI-DENSE No.
4 additive. Because this product has a finer particle size distribution, it promotes
better solids suspension and makes slurry design easier. Even so, good suspen-
sion properties still will be required to prevent settling. A weighting material that
is even finer than HI-DENSE No. 4 is MICROMAX. This product is uniquely
suited for offshore applications because it can be added directly to the mixing
water where it will remain in suspension; both HI-DENSE materials must be
dry-blended with the cement. Densifying the slurry, by lowering the water-to-
cement ratio, also can improve solids suspension.

SSA-2 also has been known to cause settling problems in high-temperature


cement designs because it has a reasonably large particle size. If slurry segrega-
tion is a problem, SSA-1 can be substituted for some or all of the SSA-2 in the
design. This modification will result in improved solids suspension while still
providing good strength stability.

When a cement design has settling problems that result in upward migration of
free fluid, the low shear rheology, yield point, and/or static gel strength develop-
ment of the slurry must be increased. Many different additives are available. Any
material that viscosifies a cement design will improve the capability of the
cement to suspend solids and prevent fluid from migrating up through the slurry.
Additives such as Microblock, Silicalite 97L, and GasCon 469 can be used.

When any of these additives are used to aid suspension in high-temperature


applications, the cement slurry can be viscous at room temperature, yet still have

Figure 1.11Example of a gas channel.

1-12 High-Pressure/High-Temperature Cementing April 1996


segregation problems under downhole conditions. To combat the problem,
Halliburton developed SA-541, a suspending aid and free-water control agent
that does not affect surface mixing viscosity. This additive yields downhole to
counter thermal thinning of cement slurries. As a result, the viscosity profile of
the slurry should remain relatively flat and uninfluenced by temperature. SA-
541 suspending aid will provide solids suspension at temperatures over 400F.

Gas Migration Theory and Control


Halliburton Research has studied extensively how gas migration causes cement
to either keep gas confined to its necessary production pathway or to remain
within its zone. Two primary types of gas migration have been identified: short-
term and long-term gas migration. Short-term migration occurs before the
cement sets, and long-term migration develops after the cement has set (Figure
1-11). The following sections provide updated research on gas migration theory
and describe control methods for both types of gas migration.

Causes of Short-Term Gas Migration


Of the two types of migration, short-term gas migration has been studied more.
The most widely accepted cause for gas migration through unset cement is that
the cement is incapable of maintaining overbalance pressure (Figure 1.12). After
the cement slurry is placed downhole, it initially behaves as a fluid and fully
transmits hydrostatic pressure to the gas-bearing formation. This overbalance
pressure prevents gas from percolating through the cement slurry (Figure 1.13-A
on Page 1-14).

Cement Cement Cement Cement


Fluid Gels Sets Hardens

A
Hydrostatic Pressure

Overbalance
Pressure Formation Gas Pressure
B

Time

Figure 1.12Maintaining overbalance pressure.

April 1996 Slurry Design 1-13


Some time after the slurry is placed in the annulus, it will develop gel strength
(Figure 1.13-B on the Page 1-14). Gelation allows the cement to support its own
weight, reducing the capability of the column to transmit hydrostatic pressure to
the gas zone. As gelation occurs, the cement loses filtrate to permeable forma-
tions, causing a loss of overbalance pressure, which then allows gas to enter the
annulus and percolate through the gelled cement (Figure 1.13-C).

A B

Permeable Filtrate
Zone Loss

Filtrate
Gas Zone Loss

C D

Filtrate
Loss

Gas
Channel

Figure 1.13A gas channel forms under the following conditions:


A. Initially after cement placement, slurry behaves as a fluid and transmits full hydrostatic pressure.
B. Static gel strength development begins; meanwhile, fluid is lost from cement slurry to permeable
formations, causing volume reductions.
C. Cement slurry static gel strength reduces transmission of hydrostatic pressure simultaneously as
volume losses occur. Together these factors cause loss of overbalance pressure, permitting gas to
enter and percolate through the unset cement.
D. Gas percolation leads to the formation of discrete gas channels throughout the unset cement. Gas
may channel to a lower pressure zone or back to the surface. Once formed, these channels will
remain in the set cement.

1-14 High-Pressure/High-Temperature Cementing April 1996


If gas begins to migrate, it will continue to percolate at a rate proportional to the
volume reductions occurring in the slurry until the cement has developed
enough gel strength to prevent further percolation (Figure 1.13-D).

Static Gel Strength


In a true fluid system, hydrostatic pressure is present. The potential loss of
hydrostatic pressure transmission is directly proportional to the level of static gel
strength (SGS) development, which is a dynamic, progressive physical property.
The length and diameter of the cement column also affects hydrostatic pressure
loss. The relationship between expected maximum pressure restriction and SGS
development can be expressed through the following equation:

SGS L
MPR =
300 D

where

MPR = theoretical maximum pressure restriction, psi


SGS = static gel strength, lb/100 ft2
300 = conversion factor (to obtain MPR in psi), lb/in.
L = length of cement column, ft
D = effective diameter of cement column, in.
(hole diameter minus pipe diameter)

In this case, MPR is a change in hydrostatic pressure that results from the
development of static gel strength.

Not all SGS development is detrimental. A certain level of static gel strength
development can prevent gas from percolating through the unset cement matrix.
The exact SGS level is not known; however, laboratory and field results show
that a 500 lb/100 ft2 SGS can prevent gas percolation or channeling through
unset cement. Therefore, gas migration will not occur if the hydrostatic pressure
is higher than the pressure of the gas-bearing formation when the cement
reaches 500 lb/100 ft2 SGS.

If the hydrostatic pressure falls below the formation pressure before this SGS
develops, gas will begin to percolate through the unset cement matrix, forming a
channel. Once a flow channel has developed, there is no level of gel strength that
can cause that channel to heal. The channel is permanent and can be removed
only by remedial (squeeze) cementing.

Transition Time
Cement slurries undergo a phase transformation from liquid to solid after
placement. During this transformation, the cement behaves neither as a solid nor
as a fluid but retains some properties of each. In this stage, the SGS of the
cement slurry steadily increases. This gel strength results from the start of
hydration. The first measurable SGS development occurs as the slurry starts the

April 1996 Slurry Design 1-15


transition from a true hydraulic fluid, capable of transmitting full hydraulic
loads, to a solid having compressive strength. The point at which the slurry loses
the capability to fully transmit the hydrostatic pressure is referred to as the start
of transition time. Throughout the rest of the transition time, the slurry will
continue to gain SGS.

This gelling process can be reversed. If enough shear (pump)


pressure is applied soon enough, the cement will revert back
7,000 psi 10,000 psi
to a true fluid capable of transmitting hydrostatic pressure. If
the slurry is allowed to remain static, it will eventually
become rigid and the gel strength will become so high that
gas cannot percolate up through the cement column. This
process is called the end of the transition time. The transi-
tion time is the time interval between the development of the
first measurable SGS and the point where the cement slurry is
so rigid a new gas channel cannot form.

Cement Slurry Volume Reductions


Reduced cement slurry volume also causes hydrostatic
A loss of 10 cc from a pressure loss. Hydrostatic pressure remains constant in a true
1,000 cc container will
cause a 3,000 psi fluid system where no fluid loss occurs. However, cement
pressure drop. slurries do not behave as a true fluid; instead, they develop
SGS before setting, preventing full transmission of hydro-
static pressure. Any fluid loss from the fluid system during
the transitional period causes a corresponding loss in hydro-
static pressure. This pressure loss can be substantial enough
Figure 1.14Fluid loss and pressure drop. to cause complete loss of overbalance pressure (Figure 1.14).
Fluid-loss additives limit the rate and volume of fluid loss
from a cement slurry, thereby limiting hydrostatic pressure
losses caused by slurry volume reductions.

Gas-Flow Potential
The gas-flow potential (GFP) factor is used to determine the most effective gas
migration control cementing system. The system should produce effective
control at the least expense to the customer without technological overkill. The
following equation shows the gas-flow potential factor.

GFP = MPR/OBP

where

GFP = gas-flow potential factor


MPR = 1.67 LD (maximum pressure loss possible at 500lb/100 ft2 static
gel strength value), psi
OBP = overbalance pressure (hydrostatic pressure minus the formation
pressure), psi

1-16 High-Pressure/High-Temperature Cementing April 1996


Gas Flow Potential Factor

1 2 3 4 5 6 7 8 9 10

Flow Condition 1 Flow Condition 2 Flow Condition 3


Minor Moderate Severe

Figure 1.15Gas flow potential (GFP) scale.

GFP is a dimensionless number indicating the estimated severity or potential for


encountering gas migration. This equation uses a static gel strength value of 500
lb/100 ft2 because SGS of this magnitude will not permit gas percolation.

Gas Migration Control Systems


Halliburtons GASFLO analysis system is a hands-on, interactive analysis tool
for modeling downhole conditions. For any gas flow situation, this program can
evaluate effective gas migration control techniques (such as adjusting cement
column length to increase hydrostatic pressure) by using the gas-flow potential
factor. By employing these simple design factors, it is possible to reduce flow
potential at little or no added expense.

Minor Gas-Flow Potential Conditions


When a low probability for flow potential is indicated, it is possible to achieve
gas-migration control without using any special application additives used in
moderate or severe cases. Any method that can control extreme conditions
would be expected to control lower flow-potential conditions. Halliburtons aim,
however, is to provide effective control in the most economical manner
without technological overkill. By using fluid-loss control additives and altering
elements of job design, many minor flow conditions can be controlled. Fluid-
loss control ranging from 50 to 100 cc/30 min is usually recommended.

Moderate Gas-Flow Potential Conditions


If the well has a moderate potential for gas flow, excellent fluid-loss control is
required. Although the fluid-loss value recommended decreases as the gas flow
potential increases, a common cement slurry recommendation in high-tempera-
ture wells is to have 25 cc/30 min of fluid-loss control. For added gas-flow
prevention, these designs can be supplemented with additives (such as GasStop
additive) that delay the slurrys gel strength development. This property permits

April 1996 Slurry Design 1-17


the cement slurry to transmit hydrostatic pressure much longer than conventional
designs. By the time the cement finally begins to gel, the rate of filtrate being
lost to the formation drops to a much lower level. As a result, the pressure drop
that occurs during the critical transition period is reduced (Figure 1.16).

Gel Strengh

A = Normal Slurry
C B = Delayed Gel-
A Strength Slurry
C = Fluid-loss Rate

Time

Figure 1.16Gel-strength rate and fluid-loss rate.

Severe Gas-Flow Potential Conditions


For severe gas-flow conditions in high-temperature wells, fluid-loss control
additives, job modifications, or delayed gelling agents alone cannot sufficiently
reduce flow potential. Highly compressible GAS-CHEK or Super CBL
cements can be used to solve severe gas flow conditions. In these cement
systems, gas bubbles are generated by a chemical reaction and are thoroughly
dispersed throughout the cement column. This action creates a highly compress-
ible cement system that can compensate for volume decreases caused by filtrate
loss and hydration volume reductions (Table 1.2). The following equation shows
the effect of increasing slurry compressibility:

Table 1.2Pressure Change vs. Compressibility


Temp. = 230F, Pressure = 7,000 psi
% Gas Volume Loss Pressure Loss
0 1 6,600
2.5 1 2,150
5 1 1,275
10 1 700

1-18 High-Pressure/High-Temperature Cementing April 1996


P = V CF

where

P = pressure loss from volume reduction


V = volume reduction caused by fluid loss and cement hydration
CF = compressibility factor

The compressibility factor (CF) for standard cement slurries is the same as for
water. GAS-CHEK and foam cements have significantly higher compressibility
factors than noncompressible cement slurries. By substituting higher values for
the CF, the ratio between volume reduction M and CF can be significantly
lowered, which lowers the value of P. Relatively low gas volumes (21/2 to 5%)
can greatly increase CF and control P. Typically only 21/2 to 5% gas by volume
is required downhole to produce enough compressibility to help prevent gas
entry into the cement column. Typical GAS-CHEK cement blends have 21/ 2 to
5% gas by volume.

Long-Term Gas Migration


Long-term or delayed-onset gas leakage occurs some time after the cement job
was performed and considered successful. As with short-term gas migration, the
best method of eliminating long-term gas migration is squeeze cementing.
However, carefully designing the cement slurry, planning the job, and using
specific cement additivesparticularly expansion additivescan help prevent
long-term gas migration.

Long-term gas migration is generally indicated by gas flow at the surface

Figure 1.17Example of long-term gas migration problems.

April 1996 Slurry Design 1-19


through the annulus, sometimes as early as a few weeks after the cement job is
performed. Flow volumes are slight-to-moderate and become more severe in
time. A noise log is probably the most reliable method of locating the source of
the problem. Cement bond logs (CBL) may not be sensitive enough to detect a
discontinuity in the cement sheath, and pulse echo technology (PET) only
evaluates the cement/pipe interface. An example of the type of flow that might
occur through a microannulus is shown in Table 1.3. Based on a slot flow
formula for a 2-in. wide channel with a formation pressure of 5,000 psi across a
1,000-ft zone at 215F, the gas flow rates would be obtained for the given
channel thicknesses listed.

To calculate the approximate scf/min,


Table 1.3Gas Flow Rate multiply the gas flow rate value by 340. For
Through Microchannel the smallest channel size presented in the
table, a value of more than 8.3 x 104 scf/d
C h a n n e l S iz e G a s F l o w R a te would be obtained, while the most severe
(in . ) (S C F / m in a t 5 ,0 0 0 p si ) case presented would result in a gas flow of
0.00 5 0.17 more than 4.6 x 106 scf/d. Either case
0.01 0 0.59 represents a costly and potentially hazardous
0.02 0 1.90 situation. It is preferable to avoid long-term
0.05 0 9.40 gas migration if possible.

Causes of Long-Term Gas Migration


There are two suspected causes of long-term gas migration: inadequate drilling
fluid displacement and the cement debonding from the casing after setting.
Inadequately displacing the drilling fluid during cementing prevents a good bond
from forming between the pipe and the cement and/or the cement and the
formation. Incomplete displacement or excessive filter cake buildup also can
create drilling fluid channels in the cement. As time passes, the drilling fluid and
cake dehydrate and shrink due to gas flow, resulting in a highly permeable
pathway for gas migration.

The second suspected cause is the cement separating from the casing after it has
set. One reason for this debonding is that the casing diameter changes after the
cement has set because of pressure or temperature changes during workovers or
stimulation treatments. The resulting long-term gas migration occurs through a
discontinuity in the cement sheath either through (1) micro-flow channels in the
drilling fluid or (2) through microannuli between the pipe and cement or be-
tween the formation and cement.

When gas is flowing through drilling fluid channels and filter cake, the flow
volume can usually be expected to increase as the drilling fluid dehydrates and
shrinks. Cements also naturally undergo a minor volume reduction during the
setting process. The magnitude of this volume reduction increases further
through fluid loss from the cement slurry, which in turn is a function of cement
slurry composition. For these reasons, fluid-loss values should be set at low but
realistic levels to prevent excessive volume reductions. Operators should also
pay close attention to obtaining the highest drilling fluid displacement efficiency
possible.

1-20 High-Pressure/High-Temperature Cementing April 1996


Solutions to Long-Term Gas Migration
Long-term gas migration can be prevented by the following two methods:
focusing on improving drilling fluid displacement principles and/or using
expanding cement compositions. Often, improving fluid displacement efficiency
has helped to stop gas migration. Drilling fluid conditioning, pipe centralization,
and the correct application of spacers/ flushes will achieve better displacement
efficiency. Pipe movement with scratchers/wallcleaners attached and high pump
rates are also helpful.

If long-term migration is still known to occur after all these methods are used,
expansive cement compositions can help correct the conditions that cause long-
term gas migration. Expansive cements have been used successfully in oilfield
operations to produce better cement bonding that has been reported to help
control annular flow, reduce water/oil ratios, and increase casing life by mini-
mizing corrosion from well brines. The two general types of cement expansions
are plastic state expansion, which occurs before the cement completes its initial
set, and chemical expansion, which occurs after initial set.

Plastic state expansion additives such as Halliburtons SUPER CBL blend


provide expansion by internally generating a chemical reaction that creates a
gaseous dispersion throughout the cement matrix before the cement sets.

A recently developed chemical expansion additive that also can be used in high-
temperature environments is MicroBond HT cement-expanding additive. This
product has been used to obtain expansion for better CBLs and effective control
of gas leakage. The key to the effectiveness of MicroBond HT is its capacity to
react after the cement hydrates by initiating the growth of crystalline materials.
This property, in turn, provides an expansion within the cement matrix as the

Solid
Particles

Reaction Product Crystals

Figure 1.18Chemical cement expansion.

April 1996 Slurry Design 1-21


crystals form and grow (Figure 1.18). This property gives the cement the capa-
bility to heal microannuli, even after the cement has set because chemical
expansion can continue to occur months after placement. The utility of this
concept has been confirmed in field applications. Several wells in a field with
long-term gas leakage problems were cemented using the MicroBond HT
additive. No leakage from these wells has been reported to date.

Cement Job Simulation


Predicting friction pressures caused by cement is important to determine safe
pump rates that can help prevent breakdown of weak formations during cement-
ing operations. Halliburtons cement job simulator service, CJOBSIM, allows
customers to see the effects of various cement job design parameters before the
job is performed to (1) help identify potential problems before pumping starts
and (2) allow operators to make appropriate modifications. (Figures 1-19
through 1-21)

CJOBSIM can be used to improve pump rates for maximum drilling fluid
displacement efficiency by designing the highest allowable pump rates, without
exceeding the fracture gradient. It also can predict circulating pressures at any
specific time during the jobeven during free-fallwhen the well is on vacuum
and the surface pressure is zero.

One of CJOBSIM's most important features is that it allows operators to evaluate


job results by comparing the prejob simulation to on-site recorded job data. This
feature allows operators to improve future designs or to analyze and pinpoint the
probable cause of a problem job.

0 7

2,000

4,000
Measured Depth ft

Fluids Pumped
6,000
Dual Spacer
Lead Cement
8,000 Tail Cement
Drilling Fluid

10,000
10,882
12,000

14,000
15,000

Plot shows calculated slurry


position as the plug lands.

Figure 1.19CJOBSIM well schematic.

1-22 High-Pressure/High-Temperature Cementing April 1996


30
Liquid Pump Rate In
Total Flow Rate Out

20
Rate bbl/min

Fluids Pumped
Dual Spacer
10 Lead Cement
Tail Cement
Drilling Fluid


0
0 400 800 1,200 1,600 2,000 2,400
Volume In bbl
Plot shows total annular return rate and corresponding pump rate vs.
liquid volume pumped into the well.

Figure 1.20CJOBSIM rates graph.

12,600
16
Fracture pressure/gradient at 15,000 ft TVD
ECD lb/gal
Circulating Pressure psi

12,200

11,800
15

Fluids Pumped
11,400 Dual Spacer
Lead Cement
Tail Cement

11,000 Drilling Fluid
0 400 800 1,200 1,600 2,000 2,400
Volume In bbl
Plot shows total annular pressure and equivalent circulating density
vs. liquid volume pumped into the well.

Figure 1.21CJOBSIM pressure and density graph.

April 1996 Slurry Design 1-23


CJOBSIM simulates jobs by examining a variable-length series of distinct time
points. After each point, an additional increment of fluid is pumped into the well.
CJOBSIM tracks circulating pressure as the pressure increases down the tubing
and decreases back up the annulus. The program adds hydrostatic pressure to
surface pressure and subtracts friction pressure from surface pressure on the way
down. During the return trip back up the annulus, the program subtracts both
friction and hydrostatic pressure. Since surface pressure is not initially known,
an iterative process is used to make the calculations.

Friction Correlations
Numerous cement job simulation programs are available that predict complete
pressure profiles for any planned displacement rate. However, the accuracy of
predicted friction pressures depends heavily on the accuracy of rheology data
and the methods used to calculate frictional pressures. Various models are
available to describe the rheological behavior of cement slurries. By nature,
however, cement slurries are quite diverse and rheologically complex. Therefore,
depending on the flow conditions, more than one model can adequately charac-
terize a cement slurry.

Recently, Halliburton researchers conducted a comprehensive study to determine


which rheological model was suitable to describe the behavior of cement (Figure
1.22). Pipe flow data for 24 discrete cement slurries ranging from extremely
thixotropic to highly dispersed and from lightweight to densified (12 to 20 lb/
gal) indicated that cement can be accurately described by the Bingham Plastic
model (Figure 1-22). In this investigation, correlations were developed to
estimate friction losses of cement slurries in laminar and turbulent flow regimes,
as well as to predict the laminar-turbulent transition.

In this study, the plastic viscosity (PV) and yield point (YP) of these slurries
were determined by pumping them at different rates through a pipe flow loop
containing four pipe sizes while simultaneously collecting data with a flow-
through rotational viscometer (Table 1.4). As a result, reasonably good correla-
tions were developed that allow operators to predict PV and YP by calculating
the flow properties using a rotational viscometer. Although the PV from both

Table 1.4Friction Pressure Calculations (psi/1,000 ft)


5 -i n . C a si n g 5 -i n . to 7 . 5 -i n . A n n u l u s
S l u r r y R a te B in g h a m B in g h a m
(b b l / m i n ) Pow er Law Pow er Law
P l a sti c P l a sti c
2 5 3 . 0 (L ) 2 7 . 3 (L ) 9 6 . 0 (L ) 5 2 . 4 (L )
8 6 1 . 5 (L ) 4 9 . 1 (L ) 1 1 3 . 8 (L ) 9 4 . 0 (L )
15 1 4 1 . 8 (T) 1 1 5 . 4 (T)
25 2 9 1 . 6 (T) 2 5 0 . 1 (T)
N o te s
L = L a m in a r fl o w re g im e
T = Tu r b u l e n t flo w r e g i m e . Th e m i n im u m flo w ra t e t o a c h i e ve t u rb u le n t flo w fo r
B in g h a m p la s t ic flu id i s 9 . 9 0 b b l/ m in fo r 5 in . c a s in g a n d 1 7 . 8 0 b b l/ m in fo r 5
in . x 7 1 / 2 in . a n n u lu s .

1-24 High-Pressure/High-Temperature Cementing April 1996


Cement Friction
Pressure
Investigation

Figure 1.22Halliburton rheology test model.

0
10

1 1/4-in. Pipe
Fanning Friction Factorf

-1
10
1-in. Pipe
3/4-in. Pipe
Newtonian

10-2

NRep (3/4 in.) = 3,200


NRep (1 in.) = 3,800
NRep (1 1/4 in.) = 4,200
10-3
2 3 4 5
10 10 10 10
Bingham Plastic Reynolds NumberNRep

Figure 1.23Friction factor curve.

April 1996 Slurry Design 1-25


Rotational Viscometer vs. Pipe Flow Data
100
Fann Yield Point, lbf/100 ft2

80
(o) = 1.591 (o)F - 2.149
60

40

20

0
0 20 40 60 80 100 120 140
Pipe Yield Point, lbf/100 ft2

Figure 1.24Yield-point graph.

pipe and viscometer data are similar, the YP of cement flowing through pipe is
approximately 1.5 times the value calculated from rotational viscometer data
(Figure 1.24). This difference can significantly affect friction pressure calcula-
tions.

A detailed discussion of this investigation is presented in SPE 19539, New


Friction Correlation for Cements from Pipe and Rotational Viscometer Data.

Rheology Correlations
Accurately determining friction pressures is complicated further by temperature
effects that change the rheological properties of the cement slurry as it is
pumped down the pipe and back up the annulus. As the temperature increases,
the rheological properties of the slurry tend to decrease. Beyond a certain
temperature, however, the PV and YP may reach a constant value. These proper-
ties determine the friction pressures that control the rate the cement slurry may
be pumped.

Therefore, if a job simulation program cannot take into account the thermal
thinning effect that occurs downhole, it will overestimate the rheological proper-

1-26 High-Pressure/High-Temperature Cementing April 1996


Connected to
Electronic Assembly
Slurry Circulation

Sleeve

Bob

Impeller

Figure 1.26HPHT rheometer.

ties in high-temperature wells. As a result, the overall effectiveness of the job


might decrease because operators pumped slower than actually required. Con-
versely, if PV and YP are known as a function of temperature, then the cement
slurry could have been pumped at a higher flow rate, improving the displace-
ment efficiency without causing any formation breakdown.

Based on this knowledge, Halliburton researchers investigated the effect of


temperature on cement slurry rheology. They designed and constructed a new
rheometer to accurately measure cement slurry rheology under HPHT conditions
(Figure 1.26). In this study, researchers tested 23 different slurries at tempera-
tures exceeding 350F.

Because of this research, CJOBSIM now has a mechanism to consider the thermal
thinning effect. CJOBSIM can recalculate friction factors that constantly change in
an actual well based on downhole rheology predictions by using the experimen-
tally developed correlations. When extrapolating to higher temperatures, these
correlations will guard against underestimating rheological properties.

During this research, Halliburton discovered that if temperature effects were


considered, pressure effects could be neglected. A detailed discussion of this

April 1996 Slurry Design 1-27


investigation is presented in SPE 20449, New Rheological Correlation for
Cement Slurries as a Function of Temperature.

Because CJOBSIM uses all the correlations that have been developed, friction
pressures and rates required to achieve turbulent flow can now be calculated
with improved accuracy regardless of well conditions. This feature is especially
important in HPHT wells where accurate prediction is critical.

Waiting-on-Cement Time
Once the slurry has been placed downhole, it is important to know how long to
wait before performing additional work on the well. To determine this waiting-
on-cement (WOC) time, operators must understand the strength requirements
needed for the cement to perform specific tasks. The following strength recom-
mendations can be used to help make this decision:

Pipe support and zonal isolation: 100 psi


Drilling out: 500 psi
Perforating
bullets: 500 psi
hollow carrier or expendable jets: 2,000 psi or greater

Whipstock plug: 2,500 psi or greater (or harder than formation)

WOC time is determined by measuring the compressive strength development of


the slurry. During this test, the slurry sample should be cured at the temperature
adjacent to the zone of interest. For example, on a liner job, it is important that
the cement placed across the liner lap is allowed to set before drilling operations
proceed. A laboratory sample of the cement should be cured at static temperature
adjacent to the liner top. The ideal instrument for determining WOC times is the
Ultrasonic Cement Analyzer (UCA).

Since dehydrated filter cake will develop more strength than a slurry that has not
lost fluid under pressure, compressive strength tests are not applicable to squeeze
jobs. Commonly, dehydrated filter cake (nodes in perforations) will develop a
strength of several thousand psi in the first 8 hours. Therefore, a waiting period
of 4 to 12 hours is generally recommended on squeeze jobs. Washing or flushing
between stages can damage squeezed zones if they are agitated or disturbed less
than 4 hours after squeezing.

1-28 High-Pressure/High-Temperature Cementing April 1996

Das könnte Ihnen auch gefallen