Sie sind auf Seite 1von 52

Chapter 18

Offshore Operations
William H. Silcox, Chevron Corp.
James A. Bodine, Chevron Corp.
Gerald E. Burns, Chevron Corp.
Carter B. Reeds,* Chevron Corp.
Donald L. Wilson, Chevron Corp.
Edward R. Sauve, Chevron Corp.

Introduction
Offshore petroleum operations emerged in the 20th century and brought new dimensions of challenge and excitement to oil exploration
and production.
When a
structure taller than a lOO-story building is launched from
a barge, or when a small city is built and placed offshore
in 2 years, those involved deserve their feelings of pride
and accomplishment.
In nearly every corner of the globe, thousands of offshore installations with payloads from 5 to 50,000 tons
are producing gas and oil today in water depths from 10
to 1,000 ft. Although subjected to winds and waves up
to hurricane intensity, earthquakes, sheet ice, severe tides
and currents, or shifting foundations, surprisingly few
structures have succumbed to the environment despite the
difficulty in predicting environmental
forces, equipment
failure, or reservoir behavior.
This chapter can only scratch the surface of offshore
operations; detailed procedures for design and construction of structures, equipment, and facilities would require
volumes. Furthermore,
such volumes would be obsolete
before they were published. Because there is no concise
reference or set of references, this chapter describes the
fundamentals of standard practice in several disciplines
and offers guidance for the selection of appropriate offshore codes of practice and technical references.

Historical Review
In 1859, Col. Edwin Drake drilled and completed the first
known oil well near a small town in Pennsylvania.
This
well, which was drilled with cable tools, started the
modern petroleum industry. Drilling methods and tools
remained in their infancy for more than 40 years, until
hydraulic rotary drilling techniques were first used to drill

*Droeascd

the Spindletop well in 1901. By then, the petroleum industry was already moving offshore.
In 1897, near Summerland, CA, H.L. Williams extended an onshore oil field into the Santa Barbara Channel
by drilling a submarine well from a pier. This first offshore well was drilled just 38 years after Col. Drakes
well. Five years later, more than 150 offshore wells were
producing oil. Production from the California piers continues even today.
From this start, offshore drilling actually turned inland
with activity in the Great Lakes, Caddo Lake in Louisiana,
and Lake Maracaibo in Venezuela. Initially, wells were
drilled from shore-connected piers and later from wooden single-well platforms. During this period of inland offshore drilling, platform technology remained basic. The
one step forward was the change from wooden platforms
to concrete structures in Lake Maracaibo.
In the late 1920s, steel production piers that extended
a quarter of a mile into the ocean at Rincon and Elwood,
CA, were built and new high-producing wells stimulated
exploration activity. In 1932, a small company called Indian Petroleum Corp. determined that there was a likely
prospect about 1/2mile from shore. Instead of building
a monumentally
long pier, they decided to build a portion of a pier with steel piles and crossmembers. Adding
a deck and barging in a derrick completed the installation. By Sept. 1932, the 60x90-ft
steel island was
completed in 38 ft of water with a 25-ft air gap. This first
open-seas offshore platform supported a standard 122-ft
steel derrick and associated rotary drilling equipment.
Successful drilling with largely unsuccessful results was
carried on intermittently on the steel island until 1939,
when the third well was completed on the pump at 40 B/D.
In Jan. 1940, a Pacific storm destroyed the steel island.
During the subsequent cleanup, divers were used for the
first time to remove well casing and to set abandonment
plugs.

18-2

Meanwhile, the first offshore field was discovered in


the Gulf of Mexico in 1938. A well was drilled to 9,000
ft off the coast of Texas in 194 1. With the start of World
War II, however, offshore activities came to a halt. Activity did not resume until 1945 when the State of Louisiana held its first offshore lease sale.
At the end of the war, surplus Navy ships and barges
became available to the oil industry. At first, Navy landing craft (LSTs) were converted into tenders to support
drilling operations on offshore platforms. By installing
mud systems and electrical generation equipment, and by
storing consumables on the tender, engineers reduced
drilling platform payloads by a factor of 10.
The development
of tender-supported
platform rigs
pointed the way toward mobile exploratory rigs that could
move on and off location, thereby eliminating the cost of
fixed drilling platforms. During the late 1940s and early
1950s, a number of mobile rigs were developed in rapid
succession.
First was the posted barge, which consisted of a submersible barge with the drilling rig mounted on steel
columns. The barge was sunk on location with the drilling rig clear of the water. Next came the submersible,
with large vertical columns that provided enough buoyancy to transport the drilling rig while floating. These rigs
were sunk on location with the drilling rig and deck remaining above water. Finally came the jackup rig. This
rig consisted of a barge hull fitted with vertical legs that
could be jacked down until they contacted the ocean floor,
thus raising the barge, which supported the drilling rig,
clear of the water. While the bottom-supported
drilling
rigs were being developed for the shallow waters of the
Gulf of Mexico, floating drilling vessels and techniques
were being developed for offshore California. There,
water depths in excess of 500 ft were found inside the
3-mile limit.
Civil and structural engineers were largely responsible for the development of submersible and jackup rigs,
but naval architects and marine engineers were called on
to convert military ships for the drilling industry. Mechanical engineers from the oil fields designed the specialized subsea and shipboard drilling equipment.
The first floating drilling vessels were converted mine
sweepers with A-frames over the side for handling pipe
and jet bits. The pipe was jetted into the ocean floor, and
core barrels were dropped through the pipe to get cores
from the bottom of the hole. Next, Navy patrol boats were
converted into drillships with over-the-side
masts and
rotary tables. The first rotary floating drilling vessel went
into service in 1953 and was capable of drilling in 400
ft of water to depths of 3,000 ft.
The adverse motion characteristics of these ship-shaped
vessels, combined with the over the side rotary table,
encouraged offshore drillers and engineers to find ways
to reduce vessel motions. In 1955, innovative drilling engineers moved the drilling rig from over the side to the
center of the ship to reduce the effects of vessel motion.
A center well, or moon pool, was installed vertically
through the hull, and the drilling rig was mounted over
it. This breakthrough led the way to modern-day drilling
vessels. Technological advances in subsea systems, vessel station-keeping systems, moored and dynamic positioning, motion compensators,
control systems, and
navigation systems have all contributed to the success of

PETROLEUM

ENGINEERING

HANDBOOK

drillships during the past 30 years. They will be discussed


in more detail later in this chapter.
While ship-shaped vessels were being developed for
California waters, a different approach to improving vessel stability was taken in Gulf of Mexico waters. The semisubmersible,
or column-stabilized
drilling vessel, was
developed by addition of buoyant hulls to a submersible
so that it could drill while floating instead of sitting on
the seafloor. These rigs exhibited superior motion characteristics and now are used extensively in such roughwater areas as the North Sea and off the east coast of
Canada.
While mobile drilling rigs were being developed into
todays sophisticated drilling systems, platform technology was keeping pace. In 1947, the first platform out
of sight of land was built off the coast of Louisiana in
20 ft of water. From then until the 1970s, the gulf coast
dominated offshore petroleum activity with the installation of more than 5,000 offshore drilling or drilling/producing
structures. During the 1970s, the North
Sea captured most of the offshore attention with the advent of huge payload requirements, and concrete gravity
structures competed with the steel template. Eighteen
concrete structures have been installed in water depths
from 240 to 540 ft with payloads up to 40,000 tons.
Meanwhile, steel-structure technology competed successfully for smaller payloads in the North Sea and
regained favor as deeper U.S. waters were explored. In
1976, Hondo,
a pile-supported two-piece jacket, was
installed in 850 ft of water off the coast of California.
In 1978, Cognac was installed in three pieces in 1,025
ft of water in the Gulf of Mexico. Single-piece structures
became viable for deeper water as launch barges and transportation technology developed. Garden Banks was installed in one piece in 680 ft of water in the Gulf of Mexico
in 1976. Cerveza,
in 935 ft, and Liguera,
in 915
ft, were installed in the gulf in 1981 and 1982. Designs
for steel jackets for up to 1,200 ft of water are in the final design stages for placement in the Santa Barbara Channel and the Gulf of Mexico.
Many other specialty structures have been installed. In
1966, a steel gravity-oil-storage
structure was placed in
service in the Gulf of Mexico. Three 500,000-bbl steel
storage domes that resemble inverted champagne glasses
were installed in the Arabian Gulf in 1969, 1971, and
1972. Buoyant articulated columns were installed in the
North Sea in the 1970s to serve as tanker mooring devices
for loading out crude oil. Tankers and drilling vessels have
been moored by various means to support gas/water/oil
separation facilities and to provide temporary oil storage.
Breast mooring and single-point mooring systems have
been installed in water depths exceeding 100 ft to accommodate a supertankers draft. A steel gravity structure
with storage capacity of 1 million bbl of oil and a deck
payload of 30,000 tons has been installed in the North
Sea as an alternative to the concrete structures. A guyed
tower was installed in 1,000 ft of water in the Gulf of
Mexico in 1983. A tension-leg platform, the commonly
favored concept for water depths of more than 1,200 ft,
was installed in 485 ft of water in the North Sea in 1984.
Each of these special-purpose structures represents an advance in ocean engineering
technology and forwardthinking business management to support untried ideas.

18-3

OFFSHORE OPERATIONS

Progress is not always the result of new ideas or concepts but often a step-by-step improvement in existing
technology. For example, the skirt pile that is currently
part of most steel deepwater structures was first implemented in 1955, but the idea had been patented in the
19th century. The North Sea gravity structure had a precedent in a gravity platform constructed offshore in California more than 30 years ago. The guyed tower was patented
before the turn of the century. The tension-leg platform
was invented during World War II as a seadrome or floating airport. Current improvements in computerized design, transportation, and installation equipment, coupled
with an ever-increasing need for new oil supplies, is the
driving force for technological advance.
During the evolution of offshore platforms, the new
ocean engineering discipline also evolved. Ocean engineers are versed in structural engineering, soil mechanics, the hydrodynamic effects of waves and currents,
structural dynamics, statistical analysis methods, and reliability analysis techniques.
The equipment, methods, and techniques for completing, producing, and maintaining wells on the ocean floor
have also undergone tremendous advancements since the
first subsea wells were completed in the late 1950s. Early
subsea Christmas trees were made up of the same conventional valves and flanges as trees for land wells. The
one concession to underwater operations was fail-safe
hydraulic actuators on remote-control valves. These early trees were usually diver-installed and connected by
Bowlines to shore. One company developed a swimming
hydraulic wrench that was fitted with television cameras
and maneuvering thrusters. This system, integrated into
the wellhead system, was successful to a degree. It was
the first attempt to eliminate divers from subsea operations. Over the past 25 years, there has been a continuous effort to reduce dependency on divers, but divers are
still a very important part of the offshore oil industry.
Complex multiwell systems have been installed on the
ocean floor. Single-well completions have been made in
1,300 ft of water. Control systems that involve hydraulic. electronic multiplex, and acoustic signal transmission
systems are now common. Unmanned, remotely operated vehicles now are being developed that will become an
integral part of the subsea completion system. Much has
been accomplished in the past 25 years, but with exploratory drilling being done in 6,500 ft of water, even more
remains to be done in this area of subsea completions.
The search for offshore oil and gas reserves has directed
the petroleum industry to the ice-covered waters of the
Arctic. In 1963, the first commercial oil field was discovered in the upper Cook Inlet of Alaska. For the first time,
ice driven by extreme tidal currents produced loads on
the production facilities far in excess of other environmental forces. By the end of 1968, 14 platforms were
producing oil and gas from the inlet.
The onshore oilfield discoveries of Prudhoe Bay in 1968
and Kuparuk in 1969 established the Alaska North Slope
as an oil province. In 1977, construction of the TransAlaska Pipeline System was completed, and oil began
flowing directly to the ice-free port of Valdez. This development has inspired extensive exploration activity in
the Arctic offshore continental shelves of the U.S. and
Canada.

Fig. 18.1 Typical floating drilling arrangement.

The industry has constructed 26 sand and gravel islands


for exploratory drilling in water depths to 100 ft since
1972. Several caisson-retaining systems have been implemented to speed construction and to reduce the fill requirements for the islands. Beyond 100 ft, drillships have
been used, but they operate only during the ice-free summer season. In 1983, a floating conical drilling unit was
deployed in the Canadian Beaufort Sea. The unit is capable of resisting early winter ice loads, hence extending
the drilling season to 6 months a year.
At the current time, at least four major Arctic marine
projects are in the planning phases: the Arctic Pilot Project in the Canadian Arctic Islands, the Arctic Marine
Hydrocarbon Production Project in the Canadian Beaufort
Sea, the Endicott Development nearshore U.S. Beaufort
Sea, and the Hibernia Development off the east coast of
Canada. Permanent production platforms, subsea pipelines, icebreaking tankers, supply vessels, and evacuation systems are a few of the facilities being developed.
In summary, though the offshore industry has come a
long way since the wooden pier days of Summerland, the
technological requirements have barely been addressed.

Offshore Drilling
The Introduction brought us quickly from the very early
days of the oil industry to todays jackup drilling units,
semisubmersibles, and drillships. This section will discuss the planning, preparation, and equipment necessary
to conduct a typical floating drilling operation (see Fig.
18.1). Focus will be primarily on floating drilling because
operations from jackups, submersibles, and platforms
generally follow land drilling practices. The last portion
of the section will be devoted to special considerations,
such as deepwater and high-current drilling and considerations for cold and hostile environmental conditions. For
a general discussion of the technology of offshore drilling, completion, and production, see Ref. 2.
Planning and Preparations
Site Conditions and Considerations. The culmination
of the sometimes arduous and complex task of geologic
evaluation of a potential offshore play is for the exploration geologist to put a finger on the map and say drill

1a-4

here. This decision sets in motion a series of actions that


will eventually lead to the drilling of an offshore well.
The first major step is to select a rig to drill the well.
Because all rigs have specific operating criteria and limits,
however, certain data must be known about the drillsite
and surrounding area. Basic rig selection criteria consist
of water depth, expected environmental
conditions during the forecasted drilling period (wind, waves, current
profile, and climatological
conditions),
distance from
nearest dock facility, and availability of consumable supplies (such as drilling mud, cement, pipe, rental tools,
and spare parts).
Water Depth. A rough idea of the water depth is an
important criterion for rig selection. If the water depth
does not exceed approximately
350 ft, any of the three
ma.jor rig types can be considered. Jackups can handle
a water depth range from their shallow draft limit of 20
to 30 ft to a maximum depth of 350 ft. The maximum
depth limitation is a function of other environmental constraints, such as wind, wave, and current conditions at
the site. Severe conditions tend to lower the jackup rigs
maximum water-depth capacity.
Drillship water depths range from approximately
100
to 8,000 ft with todays technology. The shallow side is
limited by clearance between the bottom of the hull and
the subsea blowout preventer (BOP) equipment. Maximum water-depth limits occur because of riser-system
limitations and other constraints that will be discussed
later.
Semisubmersible water depths range from approximately 150 to 8,000 ft. The semisubmersible
must stay in
slightly deeper water than a ship because of the clearance
between the submerged hull (60 to 90 ft below the water
surface during normal drilling operations) and the subsea BOP equipment. Until 1978, semisubmersible
maximum water depth was limited by the practical depth of
conventional mooring systems-approximately
2,200 ft.
One dynamically positioned semisubmersible that required
no conventional mooring system, thus extending the design working depth to 8,000 ft, was commissioned
in
1978. Today, several dynamically positioned semisubmersibles are under construction or in service.
The industry water-depth record currently stands at
6.848 ft for a well drilled off the U.S. east coast during
the summer 1983.
Expected Environmental
Conditions. Wind, waves,
and current are all important site-specific data to help in
rig selection and in determination of vessel heading, mooring pattern, mooring line tensions, riser tensions, subsea
equipment selection, and equipment operational limits.
Wind, wave, current, and climatological data are generally the responsibility of an oceanographic consulting firm
or your own companys oceanographer. Many sources of
environmental data are available-the
marine climatic atlas, ship observations,
U.S. Navy publications. privately funded oceanographic studies, and university-sponsored
research. Converting these data into useful site-specific
wind, wave, and current information
is the scientific
specialty of oceanography.
The oceanographer must have specified coordinates of
the location and the time of the year (with some cushion
on both ends) in which operations are expected. With that,
he can develop the expected wind, wave, and current conditions for the location. For an exploratory location, the
oceanographer may provide environmental data for oper-

PETROLEUM

ENGINEERING

HANDBOOK

ational weather, seasonal one-year storm, and seasonal


IO-year storm. With that information, the drilling engineer and technical support staff can accomplish several
tasks necessary in planning the well.
1. A preliminary rig selection can be made based on
water depth, wind, wave, and current information.
2. A preliminary
estimate of vessel heading can be
determined. Before a final heading is specified, however,
local knowledge of the area should be considered. Local
conditions-such
as swell, tide-generated currents, and
rapidly changing wind directions-frequently
can affect
the optimum vessel heading significantly.
The primary
objective of optimum vessel heading is to minimize vessel motion (primarily pitch, roll, and heave) while keeping the vessels mooring line forces within acceptable
limits and providing a lee side (calm-water side) for supply and crew boats to tie up.
3. To assist in vessel selection, a vessel motion or
downtime analysis can be run. Computer programs that
compare a particular vessels motion characteristics with
the predicted wind and waves are available. The result
indicates vessel motion. The resulting motion can be compared with a previously established set of acceptable operating limits (by computer analysis or manually)
to
determine an approximate downtime to be expected. This
analytical tool is most useful in comparing two rigs for
a particular location.
4. After the vessel is selected, mooring and riser analyses can be run to determine whether the vessel is adequately equipped for the location. In addition, both
mooring and riser operating tensions can be determined.
Both are necessary after the rig arrives on location. Typically, the mooring system is analyzed with a one-year
seasonal storm to determine what operating tensions
should be pulled on the anchor lines. A IO-year storm
can be analyzed to determine the level of proof test to
pull on each mooring line. With reasonable risk considered, if each line can withstand a IO-year storm proof test,
normal operations should be safe without the fear of slipping an anchor or breaking a mooring line. Drilling riser
top tensions are developed to minimize ball-joint angles
and riser sag while keeping riser-pipe stresses within acceptable limits.
For jackup rig evaluation, comparing water depth, current, wind, and tides with the maximum recommended
criteria established by the rig designer is extremely important. In water depths nearing the rigs maximum capability, strong current or other environmental
factors may
reduce the acceptable water depth.
Soil or foundation competency at the site must be known
for jackup operations also. At an exploratory location with
unknown soil consistency, soil borings generally will be
required before the rigs arrival on location. They are useful in determining depth of leg penetration and to ensure
that the soil can adequately support the rig.
Logistics Considerations. Logistics must also be considered in rig selection. Remote locations require substantially more planning and preparation than do locations
adjacent to established bases and supplies. Consideration
must be given to (1) frequency of consumables supply;
(2) distance from supply base (length of boat run); (3)
number of people the rig can accommodate; (4) availability of spare parts: and (5) shipment delays caused by
customs regulations.

OFFSHORE OPERATIONS

Floating rigs (ships and semisubmersibles) variable


deck-load capacity must be considered and compared with
frequency of consumable supplies required. Ships, as an
example, have much greater variable deck-load capacity
than semisubmersible drilling rigs (15,000 vs. 3,000 tons).
If the location is in an extremely rough environment, however, the semisubmersible is more stable in rough seas
than the ship. Trade-offs and compromises are necessary
ingredients in rig selection.
Availability of pipe, mud, fuel, water, and other consumables must be carefully determined during the planning effort. Helicopters to transport personnel and light
equipment in routine and emergency situations are a necessary part of most floating drilling operations. Those located within a few minutes of the coastline and support
bases are sometimes exceptions.
Climatological conditions have a major effect on
helicopter operations. Fog and impaired visibility conditions will ground flight operations and, depending on their
extent, can have a major effect on the resupply of consumables, transportation of crews to and from support
bases, and overall rig operations. Floating ice, low temperature, and high currents offer special considerations
that are discussed at the end of the Offshore Drilling section of this chapter.
Seismic and Other Location Studies. Preparation to
drill an exploratory location will include running and
evaluating a suite of location surveys. Site surveys generally are run by seismic companies specializing in prespud
site studies. These companies will conduct the surveys,
evaluate the data, and prepare formal reports that present the data that will be useful in selecting the exact location, in preparing the mooring plan, and in determining how the top hole will be drilled.
For exploratory drilling in federal offshore waters, the
U.S. Mineral Management Service issued a set of guidelines that require certain surveys to be performed and analyzed before it will issue a permit to drill. These
guidelines cover studies on shallow geological hazards,
culture and archaeology, and biology.
The operator or lease holder must cover a minimum
prescribed grid of traverse lines in carrying out these
studies. In addition, certain minimum instrumentation is
required to be run during the surveys. These include sparker, uniboom, sub-bottom profiler, side-scan sonar, and
fathometer for surveys of shallow geological hazards. If
the drilling equipment is to be on board a floating vessel,
no bottom sampling is required. If a bottom-setting jackup barge is to be used, then a bottom sample or core must
be obtained. Side-scan-sonar, magnetometer, and fathometer are required for cultural and archaeological surveys.
For biological surveys, box-core samples of hard-bottom
areas and ocean-floor photography or TV view of hardbottom areas are required.
The shallow-hazard surveys are required for all sites.
The grid must be at least 8,000 ft on a side, centered on
the proposed location, and surveyed on 1,000-R grid lines.
The cultural surveys need to be run only in waters of less
than 400-ft depth. The biological surveys must be run in
areas where endangered species exist or hard-bottom sediments might be disturbed. Navigation and location of the
survey grid during the water-borne surveys must be accurate to within 50 ft.

Fig. 18.2 Jackup rig.

Rig-Selection Considerations
Rig-selection criteria and rig types were discussed briefly earlier. In this section, we will discuss the differences
in four rigs that are used for offshore drilling: jackups,
submersibles, semisubmersibles, and ships. We will also
consider drilling equipment, mooring systems, and procedure manuals.
Rig types. Jackup rigs (see Fig. 18.2) consist of bargeshaped hulls with three or four (sometimes more) structural or tubular legs. Jackups must be towed to location
or loaded on specially built ships for major moves. Ship
transportation of jackups is becoming more frequent as
new special transport vessels become available. Ship
transport is considerably faster for long moves (6 to 8 vs.
2 to 3 knots) and much less risky. Loading and offloading the jackup requires a calm-water site at both ends of
the move. Once the jackup is in its jacked-up position,
drilling proceeds in a way similar to land or platform operations. However, several subtle differences should be
mentioned.
First, water conditions must be relatively calm
generally less than 6- to 7-ft waves-before the rig can
jack its hull out of the water. Major concerns are impact
and lateral loading on the legs just as it comes in contact
with the ocean floor. If the rig is rolling and pitching beyond specified limits, the jacking operation must be suspended until calmer conditions prevail. The same logic
applies when the rig is jacking down.
Second, once the rig is jacked up to working position
with a safe air gap between the ocean surface and the underside of the hull, primary concerns are lateral loading
on the legs and scouring around the leg mats caused by
current. Excessive current can cause troublesome vibration, and scouring can lead to foundation failure. Both
conditions are monitored closely, and corrective actions
are taken when necessary.

PETROLEUM ENGINEERING HANDBOOK

18-6

Fig. 18.3 Submersible rig

Third, the drilling operation is similar to a land operation after the outer casing is driven or drilled and cemented
in place. Surface BOP and conventional drilling equipment are used.
Fourth, the casing extending from the ocean floor to
the rig is a structural member and should be analyzed before installation. Wall thickness and strength of the pipe
should be specified (and will vary if a mudline suspension system is used) to ensure that it will withstand the
lateral loads of the current and the axial loads of the surface BOP and successive casing strings.

Submersible rigs (see Fig. 18.3) are limited to shallowwater drilling. Once the rig is on location and ballasted
to sit on the ocean floor, drilling operations proceed as
on a land site. Foundation considerations are as important here as in jackup operations. Logistics and supply
considerations are common to all offshore operations, so
jackups and submersibles can be just as severely hampered
by fog and bad weather as floating drilling rigs.
Semisubmersible rigs (see Fig. 18.4) evolved from submersibles. Some semisubmersibles can operate when resting on the ocean floor or in their normal semisubmerged

Fig. 18.4 Semisubmersible rig

18-7

OFFSHORE OPERATIONS

Fig. 18.5 Drillships.

position. The major advantage of a semisubmersible is


that it provides a stable floating drilling platform. Roll,
pitch, and heave are minimized because minimum structure is exposed to the water plane. The rigs main disadvantage is that variable deck-load capacity is limited by
its reserve buoyancy or the amount of watertight volume
above the water line. A semisubmersible with four 50-ftdiameter columns breaking the water plane displaces about
62 tons of seawater for each foot of displacement of the
column. An equivalent 400-ft-long by 60-ft-wide ship displaces 756 tons for each foot of hull displacement. Because semisubmersibles are sensitive to variations in deck
load, they are outfitted with extensive ballasting systems
that are capable of shifting ballast rapidly to maintain
proper trim and of deballasting or ballasting as cargo is
loaded or offloaded. The semisubmersible is highly
regarded as the year-round drilling vessel for the opensea environment because it is very stable in pitch, roll,
and heave. 3
Drillships (see Fig. 18.5) are noted for their mobility
and high storage capacity. Drillships have a definite advantage over semisubmersibles because of their size and
speed. Most drillships are designed to pass through the
major canals of the world, thereby substantially reducing the distance between oceans. The distance from the
Gulf of Mexico to the U.S. west coast by the way of the
Panama Canal is 4,500 miles. The distance around South
America to the U.S. west coast (the route a semisubmersible must travel because it is too large to pass through
the Panama Canal) is 15,000 miles. The cost of moving
the ship to the west coast is generally much less than that
of moving a semisubmersible because of time savings (less
day rate) and distance savings (less fuel). Ships generally can travel at a higher speed than a semisubmersible
(12 to 13 vs. 8 to 9 knots) for even more time savings.
As pointed out in the semisubmersible discussion, the
drillship can carry a much larger variable deck load, which
offers the advantage of less frequent resupply.

Fig. 18.6 Vessel-motion

terminology.

The very nature of drillships (long, narrow hulls with


large water planes), however, dictates their sensitivity to
sea conditions in pitch, roll, and heave. Operations can
be carried out with minimum weather downtime, however, by working drillships in protected waters at seasons
when conditions are best for open-sea drilling. Clearly,
the biggest disadvantage of a drillship working in severe
environments is its motion characteristics, especially in
pitch, roll, and heave.3
Motion Characteristics. To compare the advantage of
one drilling vessel over another, their relative motion
characteristics must be considered carefully. Vessel motions for ships and semisubmersibles can be analyzed by
determining the rigs response in the six degrees of freedom (pitch, roll, heave, yaw, surge, and sway) relative
to the uniform waves (see Fig. 18.6). All vessels should
have a set of motion-response curves. The curves generally are obtained for each rig configuration in a model
basin. Each hull shape has a unique set of curves. Roll
and heave generally control the limiting operation. With
curves like those shown in Figs. 18.7 and 18.8, vessel
motion in roll and heave can be determined for a particular set of wave data representing the drilling period. Ocean
waves represent a spectrum of wave heights and wave
periods. Computer programs are available to calculate
vessel motion by entering wave data and the rigs motion
curves. The result will be a motion history of that particular rig for a specific drilling period.
Performance Evaluation. The next step is to compare
the performance of the two rigs. One performance yardstick is the weather-related downtime the rigs will suffer
under the same environmental conditions. Downtime analysis can be particularly useful when comparing available
drilling vessels for a one-well project or a complete drilling program. While one vessel may appear to be more

18-8

PETROLEUM

ENGINEERING

HANDBOOK

20
I
.

Fig. 18.7-Vessel

response-roll.

Fig. l&8-Vessel

economical because it has a lower day rate, it may cost


more to complete the job because of weather-related
downtime.
The key to weather-related downtime is identifying the
maximum limit in degrees of roll, feet of heave, degrees
of pitch, etc., that can be tolerated during each discrete
operation of floating drilling. The maximum level may
be based on equipment operating limits, safety considerations, work efficiency, potential for damage, or other
factors. Although such a limit is seldom concise, it can
be a fair comparison to evaluate relative rig performance.
Implementing an operating limit by shutting down an operation on the rig is completely a judgment call with many
variables to be considered on the spot. Each rig should
have its own set of operating limits established from experience with the rig or from experience of the rig operating personnel. Table 18.1 is an example of limiting
vessel motions for most floating drilling operations.
With the appropriate operating limits, the percent of
the time each applies, and the rigs motion history,
weather-related
downtime can be calculated. A number
of papers have been published on downtime analysis. Various techniques (both manual and computer-aided analyses) can be applied
to calculate
weather-related
downtime. 4

TABLE

l&l-DRILLING

VESSEL

OPERATING

Heave
Limit
Operation
Anchoring, running riser,
landing BOP
Running casing, coring,
well testing
Drilling, tripping, logging
Circulate and condition mud

-- ft

LIMITS

Roll
Limit

Time
Criterion
Applied
Per Well

(deg.)

(04

1.8

10

IO
12
20

3.0
3.6
6.1

3
6
10

40
30
20

response-heave

An additional item normally not included in the motionrelated operating limits is wind. High winds frequently
result in shutdown because the rig crane cannot safely handle casing or riser. This is a valid input to the rigs overall performance
and should be included in the final
downtime comparison.
Occurrences other than severe
weather also cause operating downtime. Equipment breakdown and repair downtime (sometimes the result of severe weather, but not always) must be determined from
experience and operating history with a particular rig or
company. This increment of downtime is unpredictable
and difficult to estimate.
Mooring Systems (Stationkeeping).
Once the engineers
are satisfied that a particular rig or group of rigs is capable of handling the environment of a specified offshore
location, other equipment systems must be evaluated and
compared.
Mooring equipment provided to keep the rig on location is of major significance. Major questions to be answered regarding
mooring
equipment
include the
following: (1) is the mooring line (chain, wire, or a combination of chain and wire) strong enough to withstand
the loads during the strongest anticipated storm; (2) does
the rig have sufficient wire or chain on board or available for the water depth at the specified location; (3) do
the anchor handling or supply boats that are being considered have adequate pennant-wire-handling
equipment
on board (lengths must be greater than the water depth
and sufficiently strong to handle the 30- to 40-ton anchors
and can approach 2.5 to 3 in. diameter); (4) does the vessel have adequate instrumentation
to monitor mooringline loads; and (5) does the rig have adequate chain-locker
capacity to hold the desired amount of chain, or must part
or all of the chain be stored on supply boats? (Vessels
that dont carry their own chain have greater in-transit
deck-load capability but normally will require longer to
moor up because of the additional chain-handling
requirements.)

OFFSHORE

OPERATIONS

18-9

_I

VESSEL
OFFSET

\
A

A
400

>

AJ

SYMMETRIC

NINE LINE

SYMMETRIC

EIGHT LINE

44
SYMMETRIC

1, /

,,-

EIGHT LINE &


TEN LINE

ZERO ANGLE

30-70

Fig. 18.9-Optimum

TEN LINE

45O-90
45-90

vessel position.

EIGHT LINE

Fig. 18.10-Typical

30-60

spread mooring

EIGHT LINE

patterns,
4.:
:&!

These questions must be answered to specify an adequate mooring system properly. Mooring analysis, which
is necessary to answer several of the questions, will be
discussed later in this section.
Adequate stationkeeping (keeping the vessel within acceptable limits on the location) is a result of a properly
designed and operated mooring system. Why is stationkeeping important? Ideally, the vessel should be located
directly over the well. However, wind and current forces
can cause the vessel to take an offset downstream from
the wellhead location. Waves cause the vessel to oscillate around that offset position.
It is important to keep the vessel reasonably close to
the wellhead position for several reasons: (1) the subsea
drilling equipment can accommodate angular offsets of
up to lo, but beyond that the equipment mechanically
locks up; (2) drillpipe that is rotating in the ball joint at

the top of the BOP can cause rapid and excessive wear
if the angular offset exceeds 1 to 2 for an extended length
of time; (3) excessive vessel offset can cause increased*.
riser sag, compounding both the ball-joint offset and the
wear problems. Proper monitoring of the ball-joint angle
and adjustment of the mooring system will result in a vessel offset upstream of the current and wind that will
minimize the lower ball-joint angle. Optimum vessel offset
would yield a zero ball-joint angle (see Fig. 18.9).
There are many variations in mooring patterns. Differently shaped vessels will require different mooring patterns (see Fig. 18.10).
One criterion in mooring-system
design is that the
restoring forces should be able to withstand nearly the
same storm conditions from any direction.5 The mooring pattern is designed to fit the vessel and particular environmental conditions anticipated at the site.

18-10

PETROLEUM

ENGINEERING

HANDBOOK

r-@---i
VESSEL
I

WATER LINE
-

VESSEL
-(9f
f
I
/
@

--

0S

I
ANCHOR
LINE ,

/
/
/
D

BL

TOTAL VESSEL MOVEMENT FROM ZERO


HORlZONTAL LOAD TO SPECIFIED
HORZONTAL LOAD.

TOTALhNE

ANGLE OF LINE FROM HORIZONTAL


AT ANCHOR.

/
/
@

LEGEND
HORZONTAL FORCE AT VESSEL.

0E

VERTlCAL FORCE ON ANCHOR.

ANGLE OF ,.,NE FROM HORIZONTAL


AT VESSEL

/
NOTE

ANCHOR

Fig. 18.11-Typical

catenary

The restoring forces are generated by the niboring line.


Environmental loads acting on a vessel displace it horizontally until an equal and opposite horizontal force (restoring force) is developed by the anchor and mooring lines.
As the vessel is displaced, tension in the anchor line increases because of additional line being lifted off the ocean
floor and because the vertical component of a&nor line
tension, which increases as line is lified off-bottom, is affected by the angle in the anchopline at the vessel (see
.
Fig. 18.11).
Vertical or uplifting [orces on the anchor are zero as.
proFrly designed and
long as line-remains on botto
ways have line remainoperated mooring system sho
ing on bottom during maximum storm conditions. If all
the line comes off-bottom,+the chances of dislodging an
anchor are high.
W&h a spread mooring system, vessel excursion in
moderate weathei conditions can be restricted to 2 to 3 7%
of water depth bypulling initial operating tensions in each
line. Fig. 18.12 shows the nonlinear behavior ofborizontal force (horizontal component of line tension) and vessel disphcement
_
_ for a typical spread mooring. If the vessel

I
i
-

._-,_--;
10

20

30

DISPLACEMENT,

Fig. 18.12-Single-line
catenary
displacement.

40

50

FT.
NOTE:

horizontal

REMAINING ON BOTTOM

KS1

= PSI

X 1000

force vs. horizontal

D B E AREZERO UNTIL
BECOMES ZERO

configuration.

had two opposing mooring lines and could pull tension


on each line initially, vessel displacement could be greatly
reduced for the same environmental loads because the line
would operate in a much stiffer region of its horizon/
tal force vs. displacement curve.
Initial operating tension, however, does affect the maximum line tension that will be required in maximum storm
weather. The same environmental
loads on the vessel are
produced during maximum storm weather regardless of
the value of initial operating tension. This force must be
balanced by one or more mooring lines. This restoring
force is in addition to most of the horizontal components
of the initial tension in the line. The vessel will probably
not be displaced enough to reduce the initial tension in
the leeward lines completely. In actual operations, leeward lines can be slacked off during maximum storm
weather to reduce maximum line tension and vessel offset. In general, the higher the initial tension, the higher
the maximum line tension during maximum storm condiwill result in untions. Too little initial tension ,pwever,
acceptable
vessel offset during
operating
weather
conditions. Table 18.2 identifies desirable stationkeeping
criteria.
Dynamic positioning is another method of stationkeeping where no mooring lines are used. These systems require acoustic positioning beacons, multiple thrusters on
the vessel, and an on-board computer system and are
primarily for deepwater drilling. Dynamic positioning will
be discussed briefly in the last section, Special Considerations.
Drilling-Equipment
Considerations. Rig-selection considerations should include a review of the vessels drilling equipment. Much of the drilling equipment found on
board floating drilling vessels is identical or similar to
equipment on land drilling rigs. This discussion will be
limited to equipment unique to floating drilling.
Fig. 18.13 identifies the major components of the subsea drilling system and related shipboard ,systems. The
figtire sho& some of the components of the drilling systemthat have been developed to accommodate vessel motion and water depth. The components to be explained

OFFSHORE

18-11

OPERATIONS

TABLE

18.2-DESIRABLE

CRITERIA

Operational:
Minimal weather
Maximum vessel excursion

Drilling operations can be carried out


That which results in ~3~ lower
ball-joint angle, generally
2 to 3% of water depth

Nonoperational,
But Riser-Connected:
Maximum weather condition
Maximum line tension
Minimum line remaining on bottom
Maximum vess%l excursion

Riser-Disconnected:
Weather conditions
Maximum line tension
Minimum line remaining

OF STATIONKEEPING

Seasonal l-year storm


% breaking strength
500 ft
That which results in ~5~ lower
ball-joint angle, generally
5 to 6% of water depth
Seasonal 1O-year storm
I/Z breaking strength
100 ft

on bottom

are the BOP, the flex joint, riser, riser slip joint, riser
and guideline tensioners, drillstring motion compensator,
guidelines, and control system.
BOP. The subsea BOP stack is a major change from
land or platform drilling operations. Drilling riser, extended kill and choke lines, remote hydraulic and electrohydraulic
control systems, and subsea wellhead
equipment are all product modifications needed because
the BOP was relocated on the ocean floor. The wells
major pressure-containing
components were put on the
ocean floor because of the need to compensate for vessel
motion.
A BOP stack, whether located on the surface or subsea, is considered a last resort for preventing a well kick
from becoming a blowout. Several steps are taken to control unusual well conditions before use of the well shutin device (BOP). If the previous steps have failed and it
becomes necessary to shut the well in, the shut-in equipment must be highly reliable. BOP equipment is designed
with reliability as its ultimate criterion. Because of its relative inaccessibility,
the subsea BOP requires additional
redundancy and reliability.
The BOP stack is a combination of individual BOPs
designed to shut in a well under pressure so that formation fluids that have mov&l into the wellbore can be circulated out while continuous
control of the well is
maintained.
A description of the BOP stack components is included
below (see Fig. 18.14).
Rum Preventers. The massive steel rams have rubber
seals, and are hydraulically actuated to seal off the wellbore. Pipe rams seal the annulus around the drillpipe and
are designed so that an entire string of drillpipe and collars can be suspended from a pipe joint landed on a ram.
The ram seals must be the correct size to seal; 3-in. seals
cannot be used for 5-in. drillpipe. Conventionally,
three
pipe rams are used. A fourth ram, a blind-shear ram, is
used to seal over the open hole and to shear drillpipe when
necessary: Shearing pipe is, of course, one of the last
resorts in an emergency situation. 5 Variable-bore rams
are an option that is offered$or tapered drillstrings.
Annular Preventers. Annular preventers are comprised 9
of specially designed, reinforced rubber elements that can
seal around any tubular or near-tubular objects that &ill
go through the BOPs. They will also seal over the open
hole and can pass drillpipe tool joints without severely

GUIDE LINE
TENSIONER
4 EA

4. 6, OR 6 EA

TYP

TOP

FLEX

FLEXIBLE

HOSE

CONDUCTOR

SURFACE

CASING

CASING
l

Fig. 18.13-Floating

drilling

JOINT

STORAGE

TO KILL 8 CHOKE

system.

TYP

REEL

PETROLEUM ENGINEERING HANDBOOK

18-12

Fig. 18.14 BOP stack

damaging the sealing element. Annular preventers are actuated by an annular piston that squeezes the seal into the
bore. The piston area is large relative to the other functions on the stack and, except for initial closure, should
be operated at pressures lower than the other stack functions. This decreases the possibility of extruding the rubber seal out of the preventer. 5 Frequently, two annular
preventers are used. One normally will be located above
the upper hydraulic connector so that it can be retrieved
with the riser.
Hydraulic Connectors. 3 Hydraulic connectors provide
the main pressure seal between the wellhead housing and
the BOP and between the top of the BOP and the lower
marine riser package (LMRPusually contains the top
annular preventer, flex joint, control system, and crossover to the bottom riser joint). The high-pressure wellhead housing is the male portion of the connector. It will
be a mandrel or a hub type. The connector is the female
portion and consists of a series of hydraulic cylinders that
actuate locking dogs into grooves machined into the wellhead housing or collet fingers that clamp over the wellhead housing hub. Both types of connectors use metal-ring
seals. This provides continuous metal-to-metal sealing up
through the BOP.
Kill-and-Choke Valves. These valves are the subsea
shutoff of the high-pressure kill and choke (K&C) lines
that run from the BOPs to the choke manifold on the rig.
K&C valves are hydraulically controlled from the surface and are designed to close by spring action when opening pressure is released. Some valves close hydraulically

in addition to the spring or fail-safe close feature. Two


valves in each line should always be used for redundancy. They should be located as close to the stack as possible for mechanical protection.5
Unitized BOP Stuck. The unitized BOP stack that consists of two hydraulic connectors, three or four ram
preventers, one or two annular preventers, four K&C
valves, one flex joint, and a control system is generally
handled in one or two pieces on board the rig. The complete assembly can weigh from 200,000 to 400,000 lbm
and stand 25 to 30 ft high.3
Handling and moving the BOP stack from its storage
position to the moonpool and back presents unique problems. Generally, either special overhead cranes or hydraulically actuated carts are used to move the stacks.
BOP maintenance is extremely important. The only time
available for routine maintenance is between locations.
On short field moves, this can present problems. Land
BOP systems are frequently broken down and sent to the
shop for maintenance between wells, but that is virtually
impossible to do without causing major delays on a floating drilling rig. A few rigs are equipped with backup BOP
stacks to minimize the chance of major delay.
BOP testing is done in two steps. The stack must be
completely function-tested (each of the 30 to 40 hydraulic functions actuated to verify that each works) before running. It must also be completely pressure-tested before
it leaves the deck. Each pressure-containing component
(rams, annulars, and K&C valves) must be tested to a
pressure specified by the operator. API RP 53 on BOPs6
identifies testing procedures as a minimum safe guideline.
After the BOP has been run and latched on to the subsea
wellhead, it must again be pressure-tested. Following
procedures defined by regulatory agencies, periodic function and pressure-testing must be done on the BOP equipment during the course of a well. A complete deck and
subsea BOP testing checklist simplifies frequent testing
requirements.
Flex Joints. 3 A flex joint is installed between the lower
end of the riser and the BOP stack. This joint essentially
acts as a pinned connection to minimize bending stresses
in the riser as the drilling vessel is moved by wind, wave,
and current action.
The first flex joints were made from bag-type annular
BOPs fitted over a mandrel flanged to the top of the BOP
stack. The rubber element in the preventer was inflated
against the mandrel to a pressure high enough to keep
drilling fluid in the riser from leaking past it. This type
of flex joint, which was not positively locked to the BOP,
worked fine in shallow waters (200 ft or less) where tension was not pulled on the riser.
The next flex joints were the pressure-balanced ball
joints. These joints came into existence when operations
moved into deeper waters and it became necessary to pull
tension in the riser through the ball joint into the BOP
itself. With this positive pull upward on the ball joint, it
was necessary to provide a pressurized oil pad between
the male and female halves of the ball joint to minimize
wear. Pressurized oil was provided through a line from
the surface and was contained between upper and lower
O-ring seals within the ball joint. The balancing pressure
on the ball joint was determined by dividing the tension
pulled through the ball joint by the projected horizontal
area between the ball-joint seals.

OFFSHORE

18-13

OPERATIONS

Steel-laminated elastomers now are replacing ball joints


as riser flex joints. These joints are longer-lived and require less maintenance than the pressurized ball joints.
They also eliminate the need for the pressure source and
hydraulic lines.
Some operators also require the installation of a flex
joint between the upper end of the riser and the slip joint.
Pressurized ball joints and elastomeric joints have been
used successfully in this application. Most flex joints are
designed for an angular travel of f 10 for a total included
angle of 20.
SZip Joints. 3 All floating drilling vessels, ship-shaped
or semisubmersible,
heave up and down as swells go by.
A slip joint is the link between the riser fastened to the
bottom of the ocean and the heaving drilling vessel. The
slip joint, similar in action to a trombone, consists of an
inner and an outer barrel. The outer barrel is connected
to the riser and the inner barrel to the ship. As the ship
heaves up and down, the inner barrel strokes in and out
of the outer barrel. A pair of inflatable rubber elements
mounted on the upper end of the outer barrel serve as the
seal between the barrels to prevent loss of drilling fluid.
The second seal is for redundancy.
Riser Tensioner. For a drilling riser to survive, two
things must happen.3 First, the drilling vessel must be
kept within prescribed limits as it moves about in surge
and sway. Second, the riser must be tensioned properly
so that it will not sag and ultimately be overstressed in
bending.
The controlling criterion is not vessel position relative
to the well on the ocean floor, but the angle between the
axis of the lower end of the riser and the vertical axis of
the BOP stack. This angle is called the lower riser angle.
During drilling, this angle should be kept at less than 3.
A greater angle will cause the rotating drillpipe to cut into
the flex joint and BOP stack. In extreme cases, lost circulation has resulted from a worn-through flex joint. In
normal drilling, the riser angle is kept to less than lo.
If it exceeds 3, drilling is stopped until the vessel can
be repositioned.
To keep the lower riser angle as near 0 as possible
in areas where ocean current is a factor, the drilling vessel may have to be located up-current from the well.
If the drilling vessel is located up-current, as shown in
Fig. 18.9, but inadequate tension is pulled on the riser,
the riser could sag, as denoted by the dotted line. If the
drilling vessel is moving about and there is heavy drilling fluid in the riser, the angle at the flex joint could exceed 10 and put bending stresses in the riser. If this
situation is not corrected, the riser ultimately will fail.
Hydropneumatic
tensioning units were developed to
keep constant tension pulled on the riser. Determination
of the tension required is a complex problem in which
water depth, riser size, mud weight, ocean current, vessel motion, and sea conditions must be considered. A
number of computer programs, both time and frequency
domain, have been developed to determine the tension
needed. Many oil companies that operate offshore have
their own riser programs or have access to them. These
programs give the riser tension required and the desired
vessel offset.
The tensioner system works on the principle that displacement of a relatively small amount of hydraulic fluid
against a large pressurized volume of air results in a very

Fig. 18.15-Riser

tensioner

unit.

small change in the hydraulic pressure. Variation in tension on the riser can be kept to less than 5% by proper
design.
The tensioning unit (see Fig. 18.15) consists of a series of large air storage tanks that are connected to the air
or gas side of an accumulator that serves as the interface
between the air and hydraulic systems. The tensioner is
a cylinder/piston arrangement that has wire-rope sheaves
mounted on the lower end of the cylinder and on the upper end of the piston rod that extends out of the cylinder.
A wire rope that is dead-ended on a storage reel is reeved
through the sheaves over alignment sheaves and is attached
to the outer barrel of the slip joint. As the drilling vessel
heaves up, it pulls on the line, which pulls the piston into
the cylinder, displacing fluid into the accumulator against
the large volume of air. The air is precharged to give the
desired tension. Similarly, when the vessel moves down,
the gas pressure displaces hydraulic fluid against the
piston, extending the piston rod and maintaining a constant pull on the riser.
Guideline tensioning systems, developed to keep constant tension in the guidelines, operate in much the same
manner as the riser tensioners. The only difference is that
they are smaller because less tension is required on the
guidelines.
Drillstring Motion Compensators. Without drillstring
motion compensation, 3 the drill bit would be constantly
lifting off and banging down into the bottom of the hole
as the drilling vessel heaves up and down. Weight control on the bit under these conditions without some type
of motion compensation is next to impossible. Bumper
subs (trombone-type
slip joints) in the drillstring above
the drill collars were used initially to provide some relief
from vessel motion. However, with bumper subs, once
the drillstring was in the hole, the weight on the bit (WOB)
(weight of the drill collars) was fixed and could be changed

PETROLEUM

18-14

Fig. 18.18-Drillstring

motion compensator

only by pulling the drilling assembly and changing the


number of drill collars. Another disadvantage was that
the early bumper subs were not hydraulically balanced.
Mud pressure in the drillstring that was higher than the
external pressure in the drillpipe/hole annulus, as when
jet bits are used, caused the bumper subs to pump open
and to become as stiff as the drillpipe, making them ineffective.
Balanced bumper subs that have the internal pressure
routed to both sides of the stroking member were invented to solve this problem. These solved one problem and
created another. When working in sandy drilling fluids,
the balanced bumper subs seals wore out after relatively
short runs, making it necessary to come out of the hole
with green bits to replace the worn-out, leaking subs.
Because of their short lives, the worn-out bumper subs
were repaired on board, which required an inventory of
spare parts and personnel trained in their repair.
These considerations led to development of drillstring
motion compensators (see Fig. 18.16). These hydropneumatic units are installed either between the traveling blocks
and the hook or in the crown block at the top of the derrick. These units have been successful for both drillingbit weight control and running and landing heavy subsea
equipment (such as 400,000-lbm BOP stacks). They are
common on most drilling vessels today.
Drillstring motion compensators are similar to riser tensioners in the way they function-i.e.,
a small volume of
hydraulic fluid is displaced against a large volume of pressurized gas. The weight of the drillstring is supported on
a vertical piston inside a cylinder that is connected to the
rig blocks. The piston is supported by pressurized hydraulic fluid between the piston and cylinder. As the vessel
heaves up, the piston is pulled down into the cylinder,
displacing hydraulic fluid into a gas-charged accumula-

ENGINEERING

HANDBOOK

tor. When the vessel heaves down, the piston is forced


up by the pressurized hydraulic fluid from the accumulator. The gas side of the accumulator is connected to largevolume gas bottles. The small volume of fluid displaced
by the piston against the large volume of gas gives a low
compression ratio. This means that there is very little
change in the gas pressure and the hydraulic pressure, resulting in an almost constant WOB.
At the start of drilling, the gas pressure is adjusted so
that it will barely support the weight of the drillstring.
WOB is increased simply by reducing the gas pressure.
This transfers weight from the drillstring motion compensator to the bit. As a hole is made, the blocks are lowered
to keep the compensator at midstroke of the piston. To
reduce the WOB, the gas pressure is increased. Large air
bottles are kept charged with high-pressure air for this
purpose.
Re-Entry Systems. Re-entering a 3-ft-diameter hole in
the ocean floor in shallow waters without too much current, say less than half a knot, isnt too difficult. 3 If that
same hole is put under half a mile of water in an area
with l- to 2-knot currents, the problem obviously is more
difficult.
Almost from the beginning of floating drilling, wirerope guidelines have been used to guide drillstrings,
casing, BOP stacks, and riser pipe into or onto subsea
wells. In most instances, the guidelines are anchored to
the ocean floor by the temporary guidebase. In some
cases, when the hole for the structural pile is spudded
without a temporary guidebase, the mud pumps were run
at full capacity as the bit entered the ocean bottom. This
washed a large conical hole in the ocean floor that, with
luck, could be re-entered without guidelines. However,
under these conditions, when the structural casing and the
permanent guidebase are run, the guidelines are attached
to the permanent guidebase for subsequent re-entry operations.
With the advent of dynamically positioned drillships,
guidelineless re-entry systems were developed. These systems still had temporary and permanent guidebases; however, instead of using guidelines and guideposts, they were
fitted with guidecones that provided a large target for the
tools or casing being run. TV cameras were run through
the drillpipe, casing, riser, or BOP stack (depending on
what was being run) to provide guidance into the hole or
back onto the BOP stack. Combinations of TV and sonar
also have been used for re-entry guidance. With the
dynamic-positioning
system, the driller can take control
of the drilling vessel from his station and position it as
required for re-entry . Re-entry by means of these systems
has been made in waters as deep as 6,800 ft.
Marine Risers. The first floating-drilling
systems did
not use marine risers for mud returns. 3 Hoses that were
connected below a rotating packer mounted on top of the
BOP stack brought mud returns back to the drilling vessel. The rotating packers, which sealed around the drillpipe, were very short-lived and allowed drilling fluid to
leak into the ocean when they failed. It was the failure
of rotating packers that led to development of todays marine risers.
As may be seen in Fig. 8.14, the marine riser extends
from the BOP stack on the ocean floor up to the drilling
vessel. The marine riser, in the parlance of land drilling,
is just a very long pitcher nipple. In addition to serving

OFFSHORE

OPERATIONS

as a pitcher nipple or mud-return conduit, the marine riser


serves another useful purpose: it guides the drillstring
through the BOPs and into the hole being drilled in the
ocean floor.
The riser joints can be ordered in any length desired,
but the length normally is determined
by the geometry
of the drillship. Normal riser joints are 50 ft long, and
at least one riser made of 75-ft joints is in service.
In the beginning, riser couplings were simply threaded
collars. Cross threading of couplings being made up on
a moving vessel led to the development of clamp-type couplings, piloted union-type couplings, and finally the radially driven dog-and-groove
couplings. Riser couplings
now being developed for waters in excess of 7,000 ft are
of a piloted-bolted type.
As drilling entered deeper waters and drilling vessels
ran out of space to install more and more riser tensioners, it became necessary to reduce the weight of the riser
by adding buoyancy material. Syntactic foam was used
first. Later, air cans were installed around the riser joints
to make them air buoyant. Both types of buoyancy are
now in everyday use. The cost of the dense syntactic foam
that is required for deeper waters is offset by the cost of
high-pressure air compressors for air-buoyancy
risers.
Air-buoyant risers do have one advantage over the foam
riser package: the air in the buoyancy cans can be dumped
so that the riser will plumb bob vertically below the drilling vessel and not tend to drift off with the current.
For ultradeep waters (deeper than about 10,000 ft), freestanding risers are visualized. Work done in conjunction
with the Natl. Science Foundations proposed Advanced
Ocean Drilling Program indicates that to provide the
means for rapid disconnect of the drilling vessel from the
well, it will be necessary to establish a disconnect point
in the riser at about 1,000 ft below the ocean surface. To
support the riser vertically below the disconnect point after
a disconnect, IO-ft-diameter buoyancy cans will be fitted
to an appropriate number of riser joints. The disconnect
point will include shear rams to cut the drillpipe if an
emergency disconnect becomes necessary. This intermediate disconnect point is essential because it is estimated
that pulling 10,000 ft of riser could take from 7 to 10 days,
well outside our weather-forecasting
capability.
K&C Systems. On land rigs, the K&C outlets3 on the
BOP stack are plugged directly into the K&C manifolds
on the rig floor. In floating drilling, where the BOPs may
be from several hundred to several thousand feet below
the rig floor, K&C lines must be provided to bridge the
water depth.
In the early days in relatively shallow waters, highpressure hoses were connected to the BOP stack and, as
the stack was lowered, were paid off hose reels. When
the stack was landed on the wellhead, the hoses were connected to the K&C manifolds. As water depths became
greater, the hose reels became too large for convenient
use, and another way to bridge the water depth had to
be found. This was done first by installing guide funnels
at about 15-ft intervals along the length of the riser as it
was run. These funnels were lined up with receptacles
immediately above the K&C valves on the BOP stack.
With the riser in place, screwed-pipe K&C lines were run
down through the guide funnels and stabbed into the receptacles on the stack. Their upper ends were connected into
the K&C manifolds.

18-15

This system, while functionally satisfactory, was timeconsuming to run and test, so another method was developed. This method was to make the K&C lines integral
with the riser. The tops of the K&C joints were fitted with
a female seal pocket filled with chevron packing, and the
lower end fitted with a male seal nipple. When the riser
was run, the seal nipples dropped vertically into the female seal assembly. No rotation or screwing was required.
The joints were held together by the riser couplings.
Some manufacturers of flexible high-pressure pipe now
are proposing to provide long K&C lines that would be
stored on reels and paid out with the BOP stack when it
is run down to the ocean floor. On the larger vessels now
in service, there is space for the large hose reels required.
Control Systems. The simplest way to operate an actuator in a hydraulic control system is to connect hydraulic
lines from a pressure source through control valves directly to the actuator. 3 Some actuators require two lines to
complete the control cycle; others, such as spring-return
fail-safe actuators, require only one line.
Subsea BOPs were controlled this way during the start
of floating drilling. An early stack consisting of a hydraulically actuated connector top and bottom, K&C valves,
four ram preventers, an annular preventer, and a pressurebalanced ball joint would require as many as 17 control
hoses. These hoses, bundled together, were stored on a
large hose reel. All hoses first were connected directly
to their function on the stack, then pressure- and functiontested before the stack was run. Improperly tagged hoses
led to many hours of troubleshooting
to get the stack to
work properly. This time-consuming job had to be done
each time the stack was run.
Eventually, male and female multifunction stab plates
were developed that reduced some of the hookup time,
but the same problem of larger hose reels in deeper waters
resulted. In addition, as BOPs became more complex,
as many as 30 to 40 hoses were included in the hose bundles, doubling and tripling their size. To solve the problem of large hose reels, multihose bundles, and their
slower actuator response times in deeper waters, two new
types of control systems were developed: the piloted allhydraulic control system and the direct-wired electrohydraulic control system.
Backup Control Systems. In spite of the best-laid plans
and even with two control pods providing 100% redundancy, problems or failures still occur in the most modern
control systems. It is desirable to have reliable backup
systems if the primary controls fail. This has led to development of two types of backup control systems: the
acoustic control system and the last-chance hydraulic stab
system. 3
The acoustic backup system uses acoustic signals
through the water as the control link between the drilling
vessel and the BOP stack on the ocean floor. Energy to
power the acoustic signal receiver and to position control valves is provided by dry-cell batteries. Hydraulic
energy to power selected functions on the BOP stack
comes from accumulators mounted on the BOP stack.
These accumulators are kept charged because they are part
of the normal control system. Typical functions are to
close shear rams, to close pipe rams, and to disconnect
the riser at the lower marine-riser package.
An acoustic transmitter located on a surface vessel,
drilling vessel, work boat, or other vessel is used to send

PETROLEUM

18-16

MSSEL
CHARACTERISTICS

BATtM?3RY

rl

HORIZOifIAL
FORCE KIRIZONIAL
DISPlAC'd%V

Operating

PRIVARYMSSEL

WEtiTtER
DATA

KORING REW:+t3.UATlON:

!V,
,,IU,;WMXl

f'D3RING
LINETO DC"3Y

,:i,T,,?i
OFEUTING TtUSlON
,PRCUFOR lEJ7 TEXiON

Fig. 18.17-Mooring

analysis

HANDBOOK

pennant wire); size of control hose reels (large enough


to hold additional hose; ease of installing larger ones);
size of guideline winch drums (large enough to handle
additional line); and substructure strength (enough to support the added tension requirement).
Generally, the added water depth can be accommodated, but each rig and each site should be considered
separately.

LOCATIIX?

KIIRINGLINE
CHARACTERISTICS

ENGINEERING

method.

a signal that is coded for the desired function, down to


the receiver on the BOP stack. This signal is interpreted
and the proper control valve is actuated, directing hydraulic fluid from the accumulators to the desired function.
Acoustic backup systems now are installed on most deepwater drillships. Solid-propellant gas generators also have
been tested successfully as backup subsea energy sources.
The last-chance hydraulic stab system provides the
means for actuating several selected functions when all
else has failed. A hydraulic stab that is ported to accommodate the desired functions is run down to the BOP stack
on drillpipe. It may be guided down guidelines or directed by sonar or TV. Hydraulic hoses are connected to the
stab and are run in with the drillpipe as the stab is lowered down to the receptacle on the stack. Once the stab
is in place, control is accomplished by pressuring up the
appropriate hydraulic line. The stab also can retrieve the
lower-marine-riser
package or the complete BOP stack
in the event of a failed riser. The stab receptacle is connected with shear bolts to a mounting plate on the lowermarine-riser package. The receptacle also is attached to
the lower-marine-riser
package with a heavy wire-rope
bridle. The stab contains a connector that, when lowered
into the receptacle, latches the stab to the receptacle. To
retrieve the lower-marine-riser
package, for example, the
stab is run in on drillpipe and is stabbed and latched into
the receptacle. After the lower-marine-riser-package
disconnect is actuated, the drillpipe is picked up, the shear
bolts sheared, and the load transferred to the wire-rope
bridle. The piece then is recovered by pulling the drillpipe.
Extended- Water-Depth Capability. Occasionally,
a
drilling vessel is considered that has a maximum-waterdepth capability just short of the wellsite water depth
(1,300-ft water depth with a 1,OOO-ft capacity rig as an
example). To ensure that the rig is adequate for the location, consider additional riser availability and storage
space; additional riser tension (or added buoyancy);
lengthened control hoses and TV cable; additional guideline length; mooring system adequacy (mooring lines and

Manuals

and Emergency

Procedures

Rig selection considerations


should include a review of
each drilling vessels operations manual and emergency
procedures plan. The operations manual will include drilling operations and equipment-handling
procedures. Normal operating limits for discrete drilling operations will
be specified. The emergency procedures plan should cover
detailed responses and courses of action to be followed
during marine emergencies, well emergencies, and bad
weather situations. Disconnect and hang-off procedures
must be identified, and special equipment should be on
board to accomplish the suspension under adverse conditions. An agreement on well-control procedures should
be reached between the drilling contractor and the oil company personnel. The drilling contractor personnel will implement the procedure, so if it is different from their
previous procedures, additional training should be conducted.
Mooring and Riser Analyses
Mooring Analysis. Mooring systems and the objectives
of station-keeping have been discussed briefly. The concept of the catenary and horizontal restoring force were
mentioned. Combining these forces with the wellsite water
depth, physical description of the rigs mooring equipment, and environmental
data is the task of a mooring
analysis. Several commercial computer programs are
available to perform mooring analysis. Some companies
have developed their own programs. Mooring-analysis
methods are documented in numerous articles and papers.
Two are referenced at the end of the Floating Drilling
section. In addition, API RP 2P discusses mooring
analyses.
Fig. 18.17 describes the basic procedure followed in
mooring analysis. Combining vessel characteristics and
mooring-equipment specifics with bathymetry and weather
data yields the length of mooring line to deploy, the initial operating tension, and the proof or test tension. The
results can be obtained for a number of mooring configurations to determine which is optimum or simply to verify
a recommended configuration.
Riser Analysis. Marine or drilling risers were described
earlier. Accurate performance of drilling risers can be
determined only by analysis. In floating drilling operations, the riser behaves as a string. It gains all of its structural integrity from tension. The single most important
parameter in operation of the system, therefore, is riser
top tension. Insufficient top tension can result in operational problems associated with large ball-joint angles and,
if low enough, buckling of the riser pipe body. Overtensioning, however, produces high stresses in the riser that
can result in a shortening of its life because of fatigue
cracking. For each combination of environmental conditions, mud weight, riser weight, and vessel offset, there
is an optimum range of riser top tension.

OFFSHORE

OPERATIONS

Commercial programs are available to do riser analysis. As with mooring analysis, some companies have developed their own programs. API RP 2Q addresses riser
design* and API RP 2K discusses riser use and maintenance. 9 Papers written on riser-analysis procedures are
referenced at the end of the Floating Drilling section. The
following discussion covers riser-analysis
criteria and
operational considerations, but not details of the complex
analysis.
The items considered in riser analysis are riser stress,
ball-joint angle, top tension, riser top angle, tensioner line
angle, sheave friction, and riser pipe collapse.
Riser Pipe Stress. Static and dynamic stresses in riser
pipe are calculated by the riser-analysis program. Static
loads are caused by the riser weight, mud weight, currentinduced hydrodynamic forces, the applied top tension, and
deflection of the top of the riser. Deflection of the top
of the riser is caused by vessel offset. Dynamic loads result from wave-induced water-particle motion and vessel
surge/sway motion. Wave-induced
surge/sway motion
produces dynamic riser deflections and hydrodynamic
forces because of the relative motion of the riser and the
water.
The criteria for acceptable static and dynamic stress levels is shown in Fig. 18.18. For purely static loads (no
dynamic load applied), stresses up to 50% of the pipematerial yield strength are allowed for normal operations
and stresses up to 67% of yield strength for limited or
emergency operations. These allowable stresses have factors of safety of 2.0 and 1.5, respectively. For purely dynamic stresses, the allowables have been reduced to 25 %
of the pipe-material yield strength and 25 % of the pipematerial ultimate strength because of fatigue considerations. Combined static and dynamic stress states must fall
within the recommended range indicated on the graph.
High stresses occur in the pipe-to-connector
weld and at
the base of the groove in the connector pin. These two
areas should be inspected frequently.
Bull Joint Angle. To minimize wear by the drillpipe,
the angle of approach of the riser to the BOP stack should
be kept as small as possible. Problems are minimized if
this angle is maintained to less than lo--a goal readily
attainable in a mild environment.
With moderate to severe environments, establishing an allowable ball-joint angle of 3 is a compromise between wear problems and
the application of criteria too restrictive to permit economical drilling operations.
The lower ball-joint angle is affected by many variables. Of these, rig personnel can readily adjust only riser
top tension and vessel position. The rigs riser-angle indicator should be monitored continuously and the vessel
position and/or riser tension adjusted accordingly. Changing the vessel location relative to the wellhead is the best
method of minimizing ball-joint angle. The lower balljoint (flex-joint) angle is the most important operating
criterion to maintain.
Top Tension. For long-term operations, it is not desirable to work riser-tensioner
systems at more than about
75 % of their rated capacity. To do so will result in premature failure, generally in the tensioner lines. Tension requirements can be reduced by the use of buoyancy.
Sufficient tension/buoyancy
should be specified to prevent drastic consequences should one tensioner fail. After ball-joint angle, this criterion is the most restrictive

18-17

Fig. 18.18-Recommended

stress ranges.

on tension requirements.
When operating at the recommended tension, failure of one tensioner should not cause
increases in ball-joint angle past 3) and stress should remain in the recommended range for normal operations
(see Fig. 18.18).
Minimum operating tension should always be sufficient
for emergency disconnect. An overpull at the lowermarine-riser package connector of about 50,000 lbf is recommended to ensure that the lower-marine-riser
package
and riser will retract sufficiently to clear the top of the
BOP.
Increasing riser top tension within the specified range
can reduce bottom ball-joint angle. Increased tension beyond the maximum recommended, however, will significantly increase pipe stresses and have very little effect
on decreasing ball-joint angle. At that point, the vessel
must be moved to correct excessive ball-joint angle.
Riser Top Angle. Although the lower ball-joint angle
is the most critical, the top angle must also be controlled.
Tensions selected for drilling operations include a top angle of less than 4.
Tensioner Line Angle and Sheave Friction. Variation
in tensioner line angle generally has very little effect on
riser tension. Sheave friction, however, may be substantial in some systems. If so, its effect should be compensated for in the tensioner control system so that the desired
tension is maintained on the top of the riser.
Riser Pipe Collapse. Riser-pipe material strength and
wall thickness should be sufficient to prevent collapse owing to seawater hydrostatic pressure when the riser is completely void. Reduction in collapse strength because of
axial loading and bending stress should be included. In
general, collapse considerations
become important in
water depths greater than 800 ft.
The objective of riser analysis is to specify recommended top tensions that keep the system within safe
working limits under all anticipated conditions, as described in Table 18.3.

Field Operations
With the major planning and preparation completed, we
will now discuss the sequential steps of drilling a well from
a floating vessel. The sequence of events in this description is not necessarily followed for every well drilled from
a floating vessel, but it is a method that has worked in
the past and will work in the future.

18-18

PETROLEUM

TABLE

18.3~-RECOMMENDED

Ball-Joint
Angle Range
oto

OPERATING

LIMIT

Comments

10

1 to 30

3O and increasing
5O and increasing

Maintain ball-joint angle within these


limits, if at all possible.
Maximum limit for normal operations.
Preferably should be in this range
only temporarily.
Start drillpipe hang-off procedure.
Drillpipe hung off. Start riser disconnect
procedure.

Location

The location has been plotted on the map, the seismic


work reviewed, and the drilling program written. Materials have been delivered (especially subsea wellhead
equipment, 30- and 20-in. pipe), and it is time to survey
the location. Surveys must be accurate for several reasons. Locations near lease boundaries must be accurately placed from a legal or ownership viewpoint. Well
location relative to seismic mapping is critical. If the rig
has moved off location, getting back on location requires
good survey data.
Todays techniques can provide accuracy within 10 ft.
In well-established areas-such
as around the perimeters
of the U.S., Canada, and in the North Sea-radio
triangulation systems are used. In remote areas, satellite navigation (SAT NAV) systems, with receivers located on the
floating vessels, are used. SAT NAV systems are accurate
to within 3 ft. Depending on the well location relative to
available satellites, however, it will take multiple satellite passes and approximately 24 hours to achieve that accuracy. The site may be marked with a buoy or the rig
may be surveyed indirectly. Once on location with the
mooring system set and tested, drilling is ready to begin.
See Fig. 18.19 for the sequence of operations.
Spudding

The Well

Step 1 in drilling from a floating vessel is to lower the


temporary guidebase to the ocean floor. 3 The temporary
guidebase is generally 12 by 12 ft and is outfitted with
a bulls-eye
that is observed by TV for levelness determination. This base is run on drillpipe connected to it with
a J-tool or hydraulic connector. Four wire-rope guidelines are attached to the subbase before it is lowered. The
base may be loaded with weighted rotary mud so that necessary tension can be pulled in the guidelines when drilling equipment is lowered to the ocean floor. With the base
on bottom, the drillpipe running string is marked at the
kelly bushing, averaging out the vessel heave, so that the
water depth may be determined by measuring the pipe
when it is pulled. This measurement, corrected for tide,
is the water depth from the kelly bushing to the ocean
tloor that is used in all subsequent drilling, logging,
casing, and testing operations.
Step 2 consists of running the drilling assembly-the
bit, drill collars, bumper subs, and drillpipe-down
the
guidelines, through the temporary guidebase, and into the
ocean floor to drill the hole for the structural casing. This
hole must be drilled carefully to ensure that it is kept within lo of vertical because it later will control how vertical
the BOP stack will be when it is landed on the wellhead.

HANDBOOK

Several single shots should be taken while this hole is


drilled. When drilling is completed, heavy mud is spotted in the hole to prevent sloughing or caving in. This
hole generally has a 36-in. diameter and is 100 to 200
ft deep. The structural casing probably will be 30 in. in
diameter with a 3/4-or 1-in.-thick wall. It is called structural casing because it serves as a foundation to provide
lateral support for the BOPs during subsequent drilling
operations. In addition to drilling a hole for the structural casing, it may be driven or jetted in. In these instances,
some drillers elect not to use the temporary guidebase.
Running

Establishing

ENGINEERING

30-in. Casing

In Step 3, the structural casing (with the permanent guidebase attached and the guidelines threaded through the
guideposts) is run down the guidelines into the hole. 3
Care must be taken while the 30-in. casing is run to ensure that it is tilled with water. Should this be overlooked,
it is possible to collapse the 30-in. casing. It is run on,
and cemented through, the drillpipe lowering string. Cement returns are taken on the ocean floor and may be observed on TV. To ensure a good cement job for this
critical casing string, the cement may be overdisplaced
by as much as 100%.
In Step 4, hole is drilled for the conductor casing. If
the soil the structural casing was set in is competent, a
riser is run and latched into the permanent guidebase, and
drilling returns are taken on the drilling vessel. Riser top
tension should be sufficient only to minimize lower balljoint angle. Too much top tension could result in pulling
the 30-in. casing out of the ground. If the soil is not competent, the riser is not run and returns are taken on the
ocean floor. If the riser were used with incompetent soil,
the hydrostatic head of the drilling fluid could break down
the soil at the shoe of structural casing, resulting in lost
circulation. The conductor hole may range in depth from
298 to 499 ft. The size of conductor casing commonly
used has a 20-in. OD. If the riser has been used, it is pulled
before the conductor is run.
Running

20-in. Casing

In Step 5, the 20-in. conductor is run down the guidelines on drillpipe, landed in the housing on the structural
casing, and cemented back to the ocean floor. 3 The top
of the 20-in. conductor is fitted with a 163/4- or 185/,-in.
high-pressure
wellhead housing prepared internally to
receive subsequent casing strings. The size of the casing
head is determined by the size of the BOPs on the drilling vessel. The external profile on the upper end of the
casing head is prepared to match the type of wellhead connector installed on the BOP stack. A metal-on-metal seal
ring provides the pressure seal between the connector and
wellhead.
Running

the BOP

Step 6 consists of function- and pressure-testing the BOP


stack on the deck, then running it down the guidelines
and latching it to the casing head.3 The BOPs, which
can weigh as much as 400,000 lbm and range in height
from 30 to 40 ft, normally are run on the drilling riser.
Depending on water depth, running the BOP stack can
be a short or long procedure (from a few hours to several
days). Each riser joint must be carefully inspected as it

OFFSHORE

18-19

OPERATIONS

I
iI

-LOWERING
STRING
GUIDE

I DRILL

LINES

STRING
GUIDE

POSTS
.;OWERING

RUNNING

STRING

TOOL
-PERMANENT
GUIDE

LEVELING
BULLSEYE

-FLEX

GUIDE

BASE

ARMS

JOINT

aTEMPORARY

MUDLINE.

GUIDE
1

BASE

STEP 1 LANDING TEMPORARY


GUIDE BASE

STEP 3 RUNNING PERMANENT


GUIDE BASE AND
STRUCTURAL CASING

STEP 2 DRILLING STRUCTURAL


CASING HOLE

MARINE
RISER

FLEX

JOINT

LOWER
RiSER

MARINE
PACKAGE

BLOWOU
1
PREVENTER
STACK

STEP 4 DRILLING HOLE


FOR CONDUCTOR

STEP 5 RUNNING CONDUCTOR


WITH CASING HEAD

Fig. 18.19-Floating

drilling-subsea

systems.

STEP 6 BOPS INSTALLED


READY TO DRILL
TO T D.

PETROLEUM ENGINEERING HANDBOOK

18-20

Fig. 18.20 Dynamically positioned drillship.

is run. Each riser connection must be checked for correct makeup. The total weight of the BOP is supported
by each connection. Integral K&C lines (and hydraulic
supply line if it is hard-piped) should be pressure-tested
every second or third joint to avoid an unnecessary pulling of the BOP for leak repair. In addition to careful riser
inspection, riser handling tools and the riser spider should
be inspected for cracks or damage. The riser handling tool
supports the total weight of the BOP and riser each time
another joint is added to the string. The top of the riser
is fitted with a slip joint to accommodate vessel heave and
offset. The slip joint lands in a diverter housing immediately under the rotary table. Riser tensioning lines are connected to the slip-joint barrel so that tension can be applied
to the riser when the stack is landed on the wellhead. Once
in place, the stack is function- and pressure-tested, and
all is ready to begin drilling to total depth (TD). As hole
is drilled, additional casing strings may be run through
the riser and BOPs. Periodic BOP testing after the first
casing string is landed in the wellhead housing must be
done carefully. A leaking casing-hanger-seal assembly
could collapse the casing if test pressure exceeds casingcollapse pressure. API RP 536 includes testing guidelines. If the well is to be put on production, the tubing
string also is run through the BOPs and hung off in the
casing head.
Drillstem Testing
Drillstem testing (DST) requires specific equipment not
normally installed on floating vessels. An inside-the-BOP
production tree includes redundant master valves and a
surface-actuated hydraulic latch for emergency disconnect, high-pressure piping and valves from the choke
manifold to production-equipment area, test trap with
metering equipment, storage tanks with transfer pumps,
and a flare boom. This equipment requires considerable
space. Storage-tank location may require beefing up the
local deck structure. Flare booms require special foun-

dations. Installation and operation of production well-test


equipment requires planning and significant rig-up time.
The BOP production tree space-out is critical. The tree
lands in the subsea wellhead housing, and BOP pipe rams
seal the casing tubing annulus. The tree height, including the emergency disconnect mandrel, must not extend
above the bottom of the blind shear ram. In case of emergency disconnect, the blind-shear-ram area must be clear
for shut-in.
DST is a critical operation. It must be conducted under
carefully controlled conditions. If H 2 S is possible in the
production, special precautions are necessary. Local regulations and API RP 49 cover H2.S requirements. 10
Plug and Abandonment
If the decision is made to abandon the well, it must be
plugged first. Local regulations dictate plugging details.
Abandonment plugging generally consists of laying cement plugs in the wellbore at specified intervals to just
below the ocean floor. The plugs are pressure tested as
they are installed.
Next, the subsea wellhead and bases must be recovered.
This is accomplished by cutting the casing strings approximately 15 ft below the mudline. Cutting can be done with
mechanical cutters (tools are available that can cut
133/8-in., 20-in., and 30-in. strings in one step) or with
explosives. If explosives are used, the rig may have to
be moved several hundred feet away from the wellhead,
depending on water depth, so that the explosion shock
wave wont damage the vessels hull. Retrieved wellhead
equipment and bases can be reconditioned and reused.

Special Considerations
Deepwater Drilling
Maximum water depths continue to extend. During 1983,
a well was drilled in approximately 6,800 ft of water off
the U.S. east coast. Locations in water depths greater than

OFFSHORE

18-21

OPERATIONS

_,-

. .

-I

R1

SC-in. HIGH STRENGTH


CASING ASSEMBLY

JACK-UP AT WELL SITE

Fig. 18.21-Thirty-inch

2,000 ft should be considered deep water. Mooring becomes more difficult, subsea equipment is heavier, collapse under hydrostatic
conditions
becomes critical,
equipment performance under emergency disconnect conditions, and well-control procedures require additional
considerations.
The major differences in shallow vs. deepwater floating drilling equipment are station-keeping (moored vs. dynamic positioning),
riser design (material strength vs.
buoyancy needed), BOP control systems (hydraulic vs.
multiplex), and backup systems (diver vs. unmanned).
Deepwater operations require longer calm-weather periods and improved weather forecasting to accomplish specific tasks. Running and retrieving BOPs can take just
a few hours in shallow water. In deep water, 2 or even
3 days may not be unreasonable. Relatively calm sea conditions are required during that time.
Dynamic-positioning
(DP) systems have extended
station-keeping capabilities to depths of more than 10,000
ft. DP systems consist of acoustic beacons located on the
ocean floor, hydrophones mounted on the vessels hull,
thruster units fore and aft, and an on-board computer system for control. Dynamically positioned drilling vessels
are equipped with from 12,000 to 20,000 hp for stationkeeping. Increased fuel consumption while operating in
the DP mode is a major cost increment in a rigs day rate
(see Fig. 18.20).
Cold Environment
A few special cold-weather drilling vessels are available
However, most rigs operating in cold environments
today were not designed with low-temperature
steel requirements. Highly loaded or highly stressed components
-such as the derrick substructure, lifting subs, riser running tools, riser connectors, and elevators-must
be fabricated from steels with low-temperature resistance (Charpy
impact values comparable to the temperatures encountered) if they are to function safely. Material specificatoday.

casing with helical strakes.

tions of highly loaded components should be checked and


verified before an unknown rig is taken into a lowtemperature work situation.
Other cold-weather equipment considerations include
quarters, insulation
and heating, work-area heating,
control-fluid freeze protection, water-system freeze protection, water-tank heating, and superstructure de-icing
capability.
The American Bureau of Shipping specifies requirements for ice-class rigs. Those specifications include
hull strength in the ice zone, thruster or propeller protection from ice chunks, and other specific requirements that
must be met before a rig (ship or semisubmersible)
can
be certified to work in ice areas.
High-Current

Drilling

The first concern in drilling high-current locations is stationkeeping.


Does the mooring system have adequate
strength, or the dynamic-positioning
system adequate
horsepower, to keep the drilling vessel on the location?
With the mooring analysis previously discussed or with
more sophisticated
techniques
to evaluate dynamicpositioning stationkeeping, we can determine the adequacy
of the proposed rigs stationkeeping
system.
The next concern is possible fatigue failure of 30- and
20-in. casing strings (generally in a connector) owing to
vibration. High currents can cause vibrations that induce
failure in hours under the right conditions. Any surface
current of more than 3 knots should be considered high
current. Casing strings of 30 and 20 in. have been fabricated with helical strakes to break up the vortices that
cause the severe vibrations (see Fig. 18.21).
Vortex shedding, which causes high-amplitude vibration at 90 to the direction of the current, can create severe problems in drilling risers also. Riser fairings have
been developed and used on sever,1 occasions to eliminate the troublesome vortex-shedding vibrations successfully (see Fig. 18.22). Specia! equipment, in addition to

PETROLEUM

18-22

NOTE:

strakes and fairing, can be installed on the drilling vessel


to allow successful drilling operations in high-current
areas. A floating rotary table and a moonpool risercentralizing system have been developed to accommodate
the high loads and high angles imparted on pipes and risers
during high-current periods. If current direction is fairly
consistent, installing early-warning current-meter strings
2 to 3 miles upstream from the rig can greatly assist in
coping with the oncoming current conditions.

Structures
Background

and Philosophy

As exploration and production encroaches into deeper


water and harsher environments,
the challenges of structural design increase. Environmental
load predictions,
transportation analyses, and installation procedures are
as important to understand as the more obvious structuralframe analysis. Seldom is a designer afforded the luxury
of optimizing a structure on the basis of in-place stress
analysis. More often, the transportation and installation
(lasting a few weeks out of perhaps a 20-year structure
life) will dictate the major framing patterns. Equally disgruntling to the structural designer is that most of his accomplishment
is seen by only a handful of people,
especially once it is in place. For the structures lifetime,
it is expected to support drilling and/or production operations, and the operator cares little beyond that.
So the structural engineers job can be simply defined
as getting it designed, built, and in place as quickly and
economically as possible, while ensuring functionality and
providing for minimal maintenance.
Structure

Classification

As indicated in the Historical Review, many types of offshore structures are in service. Some are better suited to
certain environmental
and operational criteria; some are
limited by availability of construction sites; and some are
chosen simply by subjective preference of an owner/operator. Selecting a structure type is the first major structural design task after environmental and operational criteria
have been defined and might require preliminary design
of several concepts before a choice is made.
Template/Jacket.
Template
was derived from the
function of the first offshore structures: to serve as a guide
for the piles. The piles, after being driven, are cut off

vortex

HANDBOOK

KT = KNOTS
D

Fig. 18.22-Riser

ENGINEERING

= DIAMETER

shedding.

above the template, and the deck is placed on top of the


piles. The template is prevented from settling by being
welded to the pile tops with a series of rings and gussets.
Hence, the template carries no load from the deck but
merely hangs from the top of the piles and provides lateral
support to them.
Some companies prefer to place packers in the bottom
of each template leg and to grout the annular space between the leg and pile from bottom to top. The structure
and piles share the axial load from the deck and the compressive and tensile loads from the overturning moment
produced by lateral wave loads. The grouted pile also provides additional strength to the tubular joints where
horizontal and diagonal bracing are welded to the legs.
Drawbacks to this system are the difficulty in ensuring
that the grout is adequately placed and of sufficient
strength to be counted in the analysis and the additional
difficulty in platform removal.
Although both top-hung and grouted structures are
loosely called templates, some prefer to call the latter a
jacket to distinguish the difference in load path. This path
is substantially different for the overturning moment as
well as axial loads. The top-hung template requires that
moment from lateral wave loads be transmitted up the
structure to be resolved into axial pile loads. The grouted jacket has a direct downward load path for shear and
moments. The novice designer would do well to learn this
distinction early in his career.
When steel structures are designed for deeper water (in
excess of 250 ft), pile-leg grouting is prevalent. Deepwater jacket designs are heavily influenced by lateral wave
loads that produce high base shear and overturning. Piles
placed through the legs of the jacket are not always sufficient to transfer these loads to the soil, so skirt piles
are added, normally in clusters around the corner legs.
This adds a new dimension to the installation procedure.
Pile guides are required up to water level, and a removable follower
must be used during pile-driving operations. Grouting procedures for the skirt sleeve-to-pile must
recognize that grout placement and inspection will be done
remotely.
Template/Jacket
Construction. A typical construction
sequence for a template or jacket calls for yard constructing the unit on a pair of skid rails, skidding the structure
onto a barge with matching skid rails, towing the barge
to location, and launching. After the structure comes to

OFFSHORE

OPERATIONS

rest, usually floating horizontally, it is upended by ballasting members at the lower end. Once upright, it is
moved onto the final site and lowered to the seabed by
continued ballasting.
To date, the maximum water depth feasible for a singlepiece jacket appears to be about 1,000 ft. The constraint
is a result of construction equipment and facility limitations, although contractors are quick to point out that capability can be extended quickly if the business climate
warrants. Meanwhile, an alternative for deeper water is
the multipiece structure. These structures are built in either
two or three pieces that are launched from separate barges.
The pieces can be mated while floating horizontally, like
the Hondo t* platform, or stacked vertically and grouted
together with pin piles, like the Cognac platform. Mating of large structural sections has proved to be expensive and involves added risk over single-piece structures,
so plans are under way to extend single-piece construction to water depths beyond 1,000 ft.
Yard facilities are tailored to build structures in the
horizontal position. Lifting equipment is huge in both capacity and reach because a deepwater structure might have
a base width of more than 200 ft, which would require
lifts to be placed to the height of a 20-story building! A
structure for 1,200 to 1,500 ft of water could have a base
width of more than 300 ft, which is beyond the reach of
todays equipment.
Since 1975, several large transportation barges have
been built with lengths of more than 600 ft and capacities
of 40,000 tons. Such a barge was used to place the singlepiece Cerveza I4 platform in 935 ft of water. Several installation contractors now have plans for super barges
with lengths of 900 to 1,000 ft and the capability of transporting jackets more than 1,500 ft long.
The increased pile loads on deepwater structures have
necessitated advancements in pile-placing technology. In
1960, driving a 48-in. pile to 300 ft penetration for an
axial capacity of 2,500 tons was a major feat. In 1984,
a designer can plan for an 84-in. pile driven to about 400
ft to develop an axial capacity of 15,000 tons. Most of
the advancement has come in the area of pile hammers.
Hammers can be built to energy levels of 1.2 million ftlbf, nearly 10 times that of the hammers used in 1960.
Also, underwater hammers are becoming more dependable, which has the advantage of eliminating pile guides
and followers for skirt piles and saving the energy normally lost through the follower.
Concrete Gravity Structures. As the name implies, these
structures have large mat foundations instead of piles and
are heavy enough to resist overturning and base shear
from lateral wave and wind loads. Whereas steel template/jacket technology was largely a product of the U.S.,
the concrete gravity structures were European designs.
Three reasons are offered for the emergence of these designs: (1) the European countries with offshore oilfields
demanded a high percentage of national content-i.e.,
design and construction money had to be spent within their
borders; (2) European design expertise, construction facilities, and construction skills leaned heavily toward concrete and away from steel; and (3) template/jacket
technology was not prepared for the huge payloads and
severe environment of the North Sea. The first major steel
structure met with severe design changes and inadequate

18-23

fabrication and transportation


facilities and was finally
placed with high cost overruns. The high cost overruns
of the first steel structures opened the door to concrete
structures. Although these structures also ran into severe
cost overruns,
they had the advantage of not being
payload-sensitive.
The concrete structures were necessarily massive to
resist overturning.
Furthermore,
the concrete members
were often controlled more from pressure than from axial forces of the deck load. Hence, changes in the payload
requirement that occurred during construction and even
for years after placement could be accommodated with
relative ease. This proved valuable as the prolific North
Sea fields were developed and facilities were continually
added to existing structures.
Most concrete gravity structures were designed to store
crude oil until it could be loaded on a transport tanker.
This was required in the early stages of development because building an adequate pipeline network for the North
Sea took many years. Only a few of the structures still
are used for oil storage, and tanker-loadingsystems
are
active only in Norwegian fields where a deep coastal
trench has delayed pipelines from coming ashore.
Concrete structures have two basic shapes. The French
design generally consists of a large, vertical, cylindrical
structure surrounded by a perforated wall. The internal
structure is capable of storing oil and supports the center
of the deck structure. Typically, the perforated wall has
6.6-ft-diameter holes at 13-ft spacing that allow seawater
to flow through as the wave passes, thereby reducing the
wave force. The wall ends about 33 ft above the water
level, and bracing extends up to support the edges of the
deck.
Gravity Platform Construction. Concrete structures are
nearly always built vertically. The lower portion of the
base is built in a graving dock. The dock is flooded to
sea level, thereby floating the base. The gate or dike is
removed and the structure is towed to a protected deepwater site in a fjord or firth. The remainder of the structure is built by use of slipform methods and increases in
draft as construction continues. If the platform has been
designed for 450 ft of water, the draft of the completed
substructure might be 300 ft.
Meanwhile, the deck is built, outfitted, and skidded onto
one or two barges, depending on the configuration.
Substructure and barge-mounted deck are towed to a mating
site where water depth exceeds the design depth. The substructure is ballasted until only a few feet are above water.
The barge-mounted deck is towed carefully over the substructure, and the substructure is deballasted until it picks
the deck off the barges.
After the mating is complete, the platform is deballasted to minimum draft for towing to location. Minimum
draft is still deep-perhaps
350 ft-and is limited by platform stability; further deballasting would raise the platform until it became top-heavy and toppled over.
The tow to site is nearly as critical as the mating operation. The tow route must be carefully surveyed, charted, and marked. Deviation from this route could result
in stranding the structure on a sand bar or hitting a submerged canyon wall. A fleet of the worlds largest tugboats is required to tow the platform safely. Power,
steerage, and standby tugs must all respond to the direction of the tow master.

18-24

PETROLEUM ENGINEERING HANDBOOK

gle of the tower can be restrained to 2 or 3. The Lena


platform in the Gulf of Mexico is the only guyed tower
built and in service to date. Twenty-four guylines support the structure; their diameter varies between 5 and
53/, in.
The vertical support system resembles the template in
that the tower is welded to the top of the piles and hangs
in tension. The piles carry the deck load directly and, in
addition, the vertical component of the guyline tension.
An additional, undesirable, axial load is caused by the
tilting of the tower. The piles on one side compress as
the tower tilts, and on the opposite side they stretch. The
resulting stress cannot be eliminated, only controlled. The
piles are clustered near the center of the tower to reduce
the stretching and compressing. The system can be designed successfully by limiting the tilt angle and providing sufficient pile length to absorb the compression.
As the tower tilts, the piles must bend through the tilt
angle in a relatively short distance near the mudline. The
stress can be controlled in this case by locating the guides
properly and determining the stiffness of the soil. The influence of the tilt angle on the pile stresses must be evaluated before the guyline system is sized.
Guyed Tower Construction. The guyed tower is lighter
and more slender than a jacket would be for the same
depth. Therefore, yard and transportation equipment can
handle a guyed tower for depths greater than that for a
template. Installation procedures, however, are more
complex because of the heavy guyline system and the large
piles.
After the tower is upended and set on the seafloor,
buoyancy must hold the structure upright until the guylines are installed. A derrick barge is needed for driving
piles, placing the deck, and setting equipment modules.
This barge normally relies on its own mooring system for
station keeping, but in this case, some modification will
probably be required to prevent the tangling of mooring
lines with guylines. The entire installation sequence must
be planned carefully to ensure safety of the structure, guyline system, and floating equipment.
Fig. 18.23 Tension leg platform

Once on site, ballasting will set the structure on bottom, and grouting fills any voids under the base. Offshore
construction and hook-up time can be held to a minimum
because onshore deck construction should have included
commissioning most of the systems.
Guyed Towers
The guyed tower is intended for use in deep water
perhaps 1,000 to 2,000 ft. It applies the principle of compliancy to wave forces. Wave forces are cyclic in nature, pushing a structure first in the wave direction and
then against it. By pinning the structure to the seafloor
instead of making it a fixed cantilever, the guylines allow the structure to sway back and forth with the wave
force, transmitting only small loads to the foundation. The
guyline system holds the tower upright and resists the
steady forces from wind and current. Depending on the
size and configuration of the guyline system, the tilt an-

Tension Leg Platform


The tension leg platform (TLP) is another structural system based on the compliancy principle. The platform
is composed of a deck structure and a buoyant hull that
is made up of a series of vertical cylindrical columns,
horizontal submerged pontoons, and tubular member bracing. The platform is tied to the seabed by a number of
tendons that are kept taut by excess buoyancy in the hull.
Fig. 18.23 shows a TLP designed for 1,600 ft of water.
Four vertical 52-ft-diameter columns are tied together with
28-ft-diameter pontoons, and four tendons at each corner column (16 total) tie the platform to the seafloor foundation template.
To date, the Hutton platform in the North Sea is the
only TLP in service. It was placed in only 485 ft of water,
although the TLP is considered by many as the most
promising structure for water depths from 1,000 to
4,000 ft.
Because the TLP is lightly restrained horizontally,
steady forces from wind and current will cause large excursions of about 5 to 10% of water depth. Second-order
wave forces, insignificant on fixed platforms or guyed

OFFSHORE

OPERATIONS

18-25

towers, can cause substantial steady offset and slow-drift


oscillation
at the natural surge period of 60 to 120
seconds.
The tendon system and well system on the TLP are the
most structurally critical and set the limits on horizontal
excursion. Often containing flex joints or tapered cantilevered pipe sections, these systems require careful analysis both for maximum stress and for fatigue life.
The foundation system is subjected to cyclic loads superimposed on a high tension. This system also requires
careful analysis based on the best possible geotechnical
data for the site.

cally has the least concrete or steel. Construction facilities and skills in the area are important, and often the
operators preference will prevail. Here are some guidelines, though, on the four concepts discussed.

TLP Construction. The TLP is structurally similar to the


common semisubmersible drilling structure, and most efforts to design it differently or to different standards result in extreme cost and schedule overruns. The size of
the TLP is related directly to the payload required and
the environment it will withstand. This normally means
that the displacement (buoyancy) will be between two and
five times that of a semisubmersible.
Therefore, many
shipyard facilities used for semisubmersible
construction
will be too small for a TLP.
Two distinctly different construction methods are available. The first is separate construction of the deck and
hull, requiring a mating sequence similar to that described
for a concrete gravity platform. The second method is
single-piece construction where the deck is built onto the
hull and equipment placed on the completed platform. The
first method offers reduced construction time; the second
allows for intermediate structural bracing to support the
deck because clearance is not required for a barge. The
choice between these two construction methods must be
made early in the planning stage to take advantage of the
benefits afforded by either concept.
The most difficult and uncertain part of TLP construction is the placement of the seafloor template or templates
and the piling. Again, several options are open-a single
template for the foundation and well system or separate
templates. Tendons may be attached to the template or
directly to the piles. Again, each system has advantages
and disadvantages, and all aspects of allowable tolerances,
seabed levelness, potential for settlement, and preference
for separate structural systems should be considered.

Concrete Gravity Structures. These structures require


deepwater construction sites in protected water. Norway
and England have such sites; the U.S. does not. These
structures can readily carry the equipment for up to
200,000-B/D production. The North Sea will probably
need more structures with these production rates; we hope
the U.S. will also. Depth limitation is probably around
700 ft, unless new construction methods are developed.

Special-Service

Structures

Many structures have been built for a specific location


or specific purpose and are essentially one of a kind or
perhaps one of a select few. Self-floating jackets, steel
gravity platforms, tanker-mooring
articulated towers,
mooring dolphins, and single-well caissons are some
special-service structures. Each of these concepts and the
innovative effort that went into them deserves a chapter
in a book, but unfortunately there is not room here. The
author recognizes the effort needed to learn the special
hydrodynamics,
structural dynamics, stress analysis, and
construction methods needed for a unique design. Many
unheralded innovative engineers have gone virtually unnoticed because their structure, though successful in the
purpose for which it was intended, simply was not needed anywhere else.
Structure

Selection

From the preceding discussion, it should be obvious that


the chosen structure is not always the one that theoreti-

Template/Jackets.
Around the U.S., these structures are
the norm for water depths from 10 to 1,000 ft. Most U.S.
offshore structures have a payload based on less than
50,000 B/D production. These small payloads are conducive to templates and jackets. Future technology might
push these structures to 1,500 ft of water; limitations will
be based on the structures natural period and cost.

Guyed Towers. These structures should be good for deep


water and heavy payloads. Because deckloads are carried
directly by the piles and only indirectly affect the tower
and guyline system, heavy payloads should not cause a
substantial increase in structure costs.
TLP. These structural systems should be relatively economical for deep water and light payloads. Increased payload directly affects buoyancy requirements, which in turn
directly increases waveloads and tendon and foundation
requirements.
Conversely,
deeper water requires only
longer tendons. With a design based on realistic understanding, this concept could be extended to a 4,000-ft
water depth.
It will be interesting to see how the future will bring
about changes.
Structural

Design Process

As previously indicated, there are many types of offshore


structures, but the following example will be confined to
a barge-launched jacket. The considerations
and procedures, however, have a parallel in the design and construction of any structure.
Methods for determining environmental
loads and for
analyzing structures are outlined in API RP 2A, API
Recommended
Practice for Planning,
Designing, and
Constructing Fixed Offshore Platforms. This document
is one of a series of recommended practices that are reviewed frequently and reflect the state of the art.
Fig. 18.24 shows a rather large offshore jacket for about
500 ft of water. Drilling and production facilities are
assumed to be in modules for placement on top of the module support frame.
The structure must be designed for a variety of conditions, including fabrication, transportation
on a barge,
launching, upending, placement, and operation. Each condition controls the size of some members. The members
controlled by in-service conditions may require several
separate analysis procedures to size them adequately.
Field Development Plan. To put the design process in
a time frame, Fig. 18.25 shows that portion of a possible
field-development
program that relates to the jacket,

18-26

Fig. 18.24-Jacket

members

as sized by design conditions.

recognizing that facilities design and construction are


parallel. The economic studies are to determine the viability of the program and must include a preliminary design of the jacket to determine a cost estimate. Analysis
will probably be limited to a single topside load and the
maximum storm wave, using a first guess at water depth,
wind speed, wave height, and current velocity.
Environmental
Criteria.
As the design progresses
through two or three of the cycles shown in Fig. 18.25,
more exact environmental
data will be needed, and the
oceanographers probably will be working to firm the data
up in parallel with the design. Close coordination between
the designer and the oceanographer is critical. The information must be available on time, sufficiently detailed,
and in a format compatible with the design procedures.
Storm directionality might determine the orientation of
the platform. Because of transportation requirements,
a
jacket is rarely symmetrical, so it will resist lateral wave
loading better along one axis than the other. The pattern
of well conductors is often dictated by the reservoir characteristics, however, so aligning the strong axis with the
most severe storm direction is not always possible.
In addition to maximum storm criteria, statistical wave
data and scatter diagrams are needed for fatigue, transportation,
and launch analysis, and for determining
weather windows for derrick barges that will be used for
pile-driving and lifting modules. This information might
point out a substantial seasonal variation in weather statistics that will dictate the time of year for installation. The
designer should anticipate the required magnitude and format of environmental
data and pass this requirement to
the oceanographer early in the project.

Fig. 18.25-Jacket

PETROLEUM

ENGINEERING

HANDBOOK

design-key

environmental

data inputs.

In-Situ Analyses. The in-situ static procedure in Fig.


18.26 shows the loading systems applied to the jacket.
The stiffness analysis assumes these loads to be static, and
the results must be combined with the wave data and the
natural periods of the structure to derive a dynamic amplification factor. The combination
processor
must
combine static and dynamic loads to determine member
and joint forces, moments, and displacements. These are
then converted to stresses and compared with those permitted by the design codes. If the design is not acceptable, members and joints must be resized and the process
repeated.
The linear stiffness analysis usually is adequate for the
steel framework, but must be adjusted for the pile/soil
interaction that is commonly nonlinear when extremeevent loads are applied. Various linearization techniques
or iteration procedures are used to arrive at the correct
pile displacement and rotation. This is probably the most
difficult and critical area in the analysis because improper
stiffness assumptions, either too high or too low, can result in a nonconservative
structure.
Transportation and Launch. The approximate location
where the structure is to be fabricated must be known to
estimate the tow route and duration. Also, the season must
be established so that sea-state predictions are appropriate. For Gulf of Mexico structures, tow routes normally
are short, and weather forecasts often can cover the entire duration of tow and launch. For the west coast, fabrication might be in the orient, with a tow duration of 6
to 10 weeks. Thus, early construction planning is required
for adequate

design and analysis.

OFFSHORE

18-27

OPERATIONS

Fig. 18.26-In-situ

static

Fig. 18.27-Analysis

analysis.

The analysis procedure for both tow and launch is


shown in Fig. 18.27. A wave-scatter diagram is used to
define significant wave height and period combinations
for the various sea states to be encountered.
Barge motions can be determined either by analysis or model tests.
Member accelerations,
dead weight. and tiedown reactions can then be fed into the jacket stiffness analysis to
determine stress levels for checking against design-code
allowables.
The launch analysis requires a step-by-step simulation
of the jacket tipping off the barge on the rocker arms until it is floating freely. Loads and reactions. including
hydrodynamic loads on submerged members, must be retained for a stiffness analysis
Fatigue Analysis. There are two distinct methods of fatigue analysis that are used in industry. The first and most
easily understood is the deterministic analysis (shown in
Fig. 18.28). This procedure will be described in some
depth. The second method is spectral analysis, which involves a statistical prediction of stress magnitude and cycle
distribution in each sea state and over the life of the structure. This method is gaining favor in industry but is too
complicated to treat adequately in the context of this
chapter.
The basic principle of fatigue damage is straightforward. although its accuracy is often debated. It states that
for a given stress level, the fatigue damage,
is equal
to the number of applied cycles, n. divided by the number of cycles required for failure, N, or
For
example, if a steel can withstand 2.5 x 10 cycles at 10
ksi before failure occurs and 6.1 x lo6 cycles are applied,
then the damage is
x 106)/(2.5 x 10)=0.244.
This indicates that the member is 24.4% damaged. The
common procedure for checking total damage is Miners
rule, which requires that the summation of all damage be
less than 1.0. The recommended fatigue curves and alF,,.

F,,

F,,

=(6.1

=tdN.

procedure

for tow and launch.

lowable cumulative damage can be found in API RP 2A.


The fatigue curves show stress, S, plotted against the number of cycles to failure, N. and are often called S-N curves.
If the natural period of the structure exceeds 3 seconds,
API RP 2A recommends that a dynamic amplification factor (DAF) be included in the analysis. The DAF accounts
for increased stress because of structural vibration. A
complete dynamic analysis may be performed in lieu of
using DAFs, but this can become extremely expensive.
Fig. 18.29 shows a spectral fatigue analysis that can
be compared to the deterministic analysis in Fig. 18.28.
Ref. 13 outlines this procedure.
The previous discussion covers most of the major analyses required for a platform, but dont be misled by the
brevity. Between 75 and 100 complete stress analyses
usually are required to determine stresses for in-place fatigue. Transportation
fatigue might require twice that
many. If seismic analyses are included, the total bill for
computer time might exceed $1 million for a thorough
final analysis.
Other structures will have unique design and analysis
problems. Generally, the same types of analyses arc required, but they differ in areas of dynamic response or
hydrodynamic loading. Ultimately. the extent and choice
of the analysis procedure falls back to engineering judgment, and there is absolutely no substitute for experience.

Offshore

Production

Operations

Offshore production installations can be either very similar


to or radically different from land installations. The purpose of this section is to provide a general overview that
will acquaint inexperienced persons with typical designs
and requirements for completing and producing wells offshore and describe alternative installations that are currently available to engineers and operators for meeting
production objectives under varying conditions.

PETROLEUM

18-28

ENGINEERING

EACH WA STATE

Fig. 18.28-Deterministic

Platform

procedure

analysis.

Production

Well systems and crude-oil and gas process facilities installed on platforms account for more than 99% of current offshore production capacity. A small number of
wells are completed on manmade islands. From a design
and operating standpoint, these island wells are handled
the same as platform wells. Wells that are completed subsea number fewer than 300 worldwide and will be discussed separately.

Well Completions. Except for a few innovative installations, wellheads and Christmas trees on platforms are basically the same as for land wells (Fig. 18.30). Control
valves, safety valves, and piping outlets are configured
the same and use the same or similar components. Some
of the valves probably will have pneumatic or hydraulic
actuators to facilitate remote and rapid closure in an emergency. Also, some Christmas trees may have composite
block valves instead of individual valves flanged together.
The major difference, however, between land and platform well completions is the economic incentive on platforms to reduce equipment weight wherever possible and
to minimize space requirements.
Simply put, lighter,
smaller equipment and more compact installations result
in less-expensive platforms. A good example is the use
of composite block valves to reduce Christmas-tree size
and weight. Another example is the spacing of wellheads
as close together as drilling operations will permit, with
just enough room for safe and efficient operation of tree
valves, control valves, and well-workover
equipment.
Typically, this means centerline distances of 6 to 10 ft
between wellheads.
Where only one drilling rig is on a platform, all the
wellheads usually are located in one bay. Larger platforms
that are designed to accommodate two drilling rigs may
have two well bays (one for each rig) with two or more
rows of wells in each bay.

Fig. 18.29-Spectral

fatigue analysis

HANDBOOK

EACH mnEcTloN

procedure

Process Equipment. The primary function of process


equipment, whether on a platform or on land, is to stabilize produced fluids and to prepare them for shipping or
disposal. Well production is separated into components
of oil, gas, and water (and sometimes condensate). The
separated fluids are measured and then either shipped, injected back into the reservoir,
or flared.
Differences between the process equipment (oil and gas
separators, free-water knockouts, gas scrubbers, pumps,
compressors, etc.) installed on a platform and those installed on land are minor (Fig. 18.3 1). Where possible,
consideration is given to using vessels and machinery that
are compact and lightweight-e.g.,
electric motors are
commonly used instead of gas engines for driving pumps
and compressors. Vertical clearance between decks may
impose height limitations and dictate, for example, the
use of horizontal instead of vertical separators.
There is a major difference, however, in the way equipment is packaged. If it is to be installed offshore after
placement of the platform jacket and decks, process equipment usually is built in modules at a land site. The module assemblies then are barged offshore, lifted onto the
platform, and hooked up. This significantly reduces expensive offshore installation and hookup time. In any
event, the equipment and its piping, wiring, and controls
are installed as compactly as possible. The extra engineering and fabrication costs needed to reduce deck area to
an absolute minimum are more than offset by savings in
platform structure cost.
Well Servicing and Well Workovers. On relatively small
platforms with no more than 5 to 10 wells, it is common
practice in some areas to drill all the wells before any of
them is placed on production. The drilling rig is removed
after the last well is drilled, and future well workovers
are performed with a portable workover rig or wellpulling unit. Downhole work that does not require pulling tubing (e.g., replacing safety valves, gas lift valves.
or standing valves) normally is accomplished with a wireline unit.

OFFSHORE OPERATIONS

18-29

Fig. 18.30 Platform well bay

On larger platforms with more wells, drilling and production operations generally are carried on concurrently. In this case, well workovers are performed by the
drilling rig if it is still on the platform. Depending on the
urgency of the workover and on economic considerations,
the work typically is scheduled to follow completion of
the well that is currently being drilled. Wireline repairs
can be performed without interfering with drilling operations unless the two wells involved are too close together
for safety considerations.
Even on the largest platforms, drilling rigs normally
are removed after all scheduled wells have been drilled.
Depending on the number of wells and the amount of
downhole work anticipated, a special workover rig may
be installed permanently on the platform. An economic
comparison between using a portable rig and using a permanently installed rig should be the basis for selection.
Crude-Oil Disposal. In the great majority of cases, crudeoil production is shipped from platforms by subsea
pipelines. Because most offshore producing areas involve
multiple platforms and more than one operating company, the pipelines are generally common carriers.
In the simplest of installations, where a pipeline transports only one type of crude oil from a single platform,
an optimum pipeline design and installation can be involved and expensive. Numerous factors must be evaluated, such as seawater temperature, seafloor profile and
geologic features, water currents between surface and

seafloor, hazards from commercial fishing equipment and


boat anchors, the possible need for a weight coating such
as concrete to ensure negative buoyancy, cathodic protection, corrosion-prevention coatings, water depth, the
possible need for burial beneath the seafloor, the type of
beach crossing that will best protect the pipe and also be
environmentally acceptable, the best riser to use at the
platform to afford protection from corrosion and physical damage from boats and waves, total length of pipeline, pumping rates and pressures, the need for periodic
pigging and inspection, safety shutdowns to prevent or
to minimize pollution in the event of failure or accident,
and the crude-oil properties and rheology. When crude
oil is shipped from more than one platform, a more detailed study of rates and pressures will be required, and
if crudes from different reservoirs are being pumped
through the same pipeline, a much more detailed study
of oil properties and rheology will be necessary.
Depending on pipe diameter, length, need for burial,
need for coatings and cathodic protection, water depth,
and various construction considerations, an offshore pipeline can be the single most expensive element of an offshore installation, sometimes far exceeding the cost of one
or more platforms. In the great majority of cases, however, piping is still the safest, most economical way to
transport crude-oil production to a land site.
Occasionally, an offshore oil field is too remote, production rates are too low, or the field is too short-lived
to justify a pipeline economically. The alternative is to

PETROLEUM

18-30

Fig. 18.31-Platform

ENGINEERING

HANDBOOK

deck layout for process facilities.

ship the oil by tankers. This usually requires some type


of loading system installed 1 to 2 miles from the platform,
such as a moored buoy or articulated loading tower. A
seafloor pipeline connects the loading facility to the platform, and a tanker is moored to the loading facility during the transfer of oil.
The two most important drawbacks of a tanker-loading
operation are sensitivity to weather and the need for
separate oil storage. Tanker loading is best suited to mildweather areas to minimize downtime from storms. Oil
storage requirements will depend on total field producing rates and reservoir characteristics (can the wells be
shut in for short periods without loss in productivity) as
well as tanker downtime. This has led to the development
of permanently moored storage tankers.
Gas Disposal. Disposition of gas from an offshore production site will depend on a combination of reservoir and
economic factors. If well production is primarily oil, the
gas may be handled as a byproduct and be disposed of
in the most economical way. Piping the gas to a land site
for sale and use as fuel is generally preferred if it can be
done economically. Injection back into the producing formation is a common alternative. This helps to maintain
reservoir pressure and conserves the gas for possible fu-

ture sale. In some areas, gas flaring is still acceptable,


but many countries now forbid it except for short test periods and for the disposal of small amounts of residual waste
gas.
Water Disposal. Produced water is normally cleaned so
that it may be either discharged offshore in accordance
with governmental
regulations or reinjected. In either
case, a combination of mechanical and chemical means
may be used to condition the produced water before disposal. Tankage and filtration are used to remove oil and
other contaminants from the water. Chemical treatment
is common to control bacteria and corrosion in injection
wells.
Subsea Completions
Seafloor well completions occupy a very small niche in
the offshore petroleum industry, but they attract a lot of
attention. Their primary use has been as single satellite
wells producing to a nearby platform. They are a means
of producing field extremities that cannot be reached by
directional drilling from an existing platform and where
the economics do not justify the installation of one or more
additional platforms. Some multiwell templates and piping manifolds have been installed that go beyond the satellite well concept.

OFFSHORE

18-31

OPERATIONS

The main benefit of subsea completion efforts is that


the petroleum industry and political governments now
recognize and accept them as a technically viable means
for producing offshore oil and gas wells. This, in turn,
means that future installations can be evaluated on the basis of actual field experience and realistic economic considerations. Subsea completions probably will increase in
popularity in coming years, especially if significant discoveries are made in deep water where conventional platforms are either extremely expensive or impractical.
Wet vs. Dry. Seafloor installations are made either with
the equipment protected by a dry, one-atmosphere pressure chamber (Fig. 18.32) or with the equipment exposed
to the sea environment (Figs. 18.33 and 18.34). The dry
chambers are large enough for workmen to install and to
repair valves, flanges, and control systems in a shirt-sleeve
environment. Access is by way of a diving bell lowered
from a work boat. Successful installations have been made
in water depths greater than 500 ft.
Most subsea completions are of the wet type and require varying amounts of diver intervention during installation and removal of the seafloor equipment. Minor
repairs and trouble-shooting
can sometimes be accomplished in place, but major equipment or control-system
repairs are made above water after the faulty equipment
is removed. Running tools operated from a floating drilling or workover rig are used for installation and removal
of well completion equipment in much the same way that
drilling operations are conducted when running and
retrieving a BOP stack for a subsea well. Wet well completions have been made in water depths greater than
1,300 ft, and a satellite system has been constructed in
2,500 ft of water. The latter is designed to use tethered,
remotely operated vehicles with manipulators instead of
divers for trouble-shooting
installation problems and for
assisting with repairs.
Single Satellite Wells. Functionally,
seafloor well completions are no different than wells on platforms or land.
Completions in shallow water where divers are used extensively for installing the equipment may even look like
land wells.
In deeper water, however, where diving requires expensive saturation systems for the divers, more reliance
is placed on equipment that can be installed and removed
with special tools that are run and operated from a floating drilling rig. Hydraulically actuated wellhead connectors are used instead of flanged or clamped connectors.
Tubing hangers that can be locked in place and tested remotely are used. Precise equipment systems that can be
remotely connected and disconnected and that permit personnel on the drilling rig to test and to function all of the
wellhead equipment are required. Hydraulic controls are
generally used for this work. The result is a Christmas
tree that may be 20 to 40 ft tall when combined with its
drilling bases and that may cost several times more than
a land tree.
One very important
requirement
for subsea-wellcompletion equipment is that it be totally compatible with
the drilling equipment that is used in drilling the well (Fig.
18.35). This requires extensive pre-engineering
and
preplanning between persons responsible for the drilling,
completion, and producing operations.

Fig . 18.32~Subsea

tree-dry

type with diving bell in place.

PETROLEUM ENGINEERING HANDBOOK

18-32

Fig. 18.33 Subsea treewet type.

Multiwell Templates. A seafloor template with guide


posts and wellhead receptacles for more than one well is
suitable for drilling a number of satellite wells in close
proximity to one another. Minor cost savings are realized during drilling operations because the drilling vessel does not have to be moved and reanchored between
wells. Any slight adjustment in position can be accomplished by taking in or letting out opposing anchor lines.
A major savings in flowlines may be possible by commingling well production at the template and transferring
it through one large pipeline instead of individual
flowlines. Well testing can still be accomplished by installing one separate flowline with a valve manifold for
switching wells. Multiwell templates offer additional opportunities for reduced costs for control systems, gas-lift
piping, and water-injection piping. Obviously, potential
cost savings will be the greatest with a large number of
wells. Multiwell template designs should consider poten-

Fig. 18.34 Exploded view of diverless subsea tree and running


tools.

tial hazards from accidentally dropped tool joints or other


heavy equipment during drilling operations. This may be
a deterrent to producing completed wells until all wells
in the template have been drilled. For a large template
with many wells, this would have a significant effect on
cash flow. A detailed economic analysis of all relevant
factors is essential to determine the optimum size and configuration of a multiwell template.
Manifolds. One technique for combining some of the advantages of single satellite wells with the piping savings
for multiwell templates is to produce moderately spaced
satellite wells to a central, seafloor manifold installation
(Fig. 18.36). The manifold would include valves and controls to commingle or test each well selectively and would
reduce overall piping- and control-system costs if platform process facilities are located several miles away.
Both dry and wet manifolds have been successfully installed and operated.

OFFSHORE

OPERATIONS

18-33

Fig. 18.35-Sequence
drawings for drilling and completing
subsea well and for well re-entry workover.

Flowlines and Control Lines. Subsea satellite wells may


be installed with either one or two flowlines, depending
on well conditions and operating requirements. The production flowline is usually 2 to 6 in. in diameter. Size
is dictated primarily by flow rate, flowline length, and
wellhead pressure. A second flowline frequently is installed for communicating
with the tubing/casing
annulus. This line is effective for monitoring annulus pressure
and for circulating kill fluid if needed. It can be hooked
up to pump pigs, paraffin scrapers, or through flowline
tools. It can also be used as a second production line.
In some cases, a decision to bury flowlines below the

Fig. 18.38-Seafloor

manifold

seafloor can be made because of local regulations or obvious hazards. Generally, however, the pros and cons for
burial and the overall costs and economics should be
evaluated carefully before a decision is made to bury
flowlines. If the answer is not clear-cut, leaving the lines
unburied is probably best. Unburied lines cost less to install, are less expensive to repair, and are easier to inspect for leaks and damage.
Conventional
welded steel pipe is used for most
flowlines. It can be protected against external corrosion
by either anodes, a corrosion coating, or a combination
of both. Cathodic-protection
methods should be compat-

for satellite

subsea wells.

18-34

ible with corrosion-prevention


designs for equipment at
both ends of the flowline (wellhead equipment and usually
a platform jacket), as well as other nearby or crossing
pipelines.
In recent years, flexible pipe made of laminations of
steel wires and other materials has become popular for
flowline service. Although the material cost for flexible
pipe is usually very high in comparison with conventional pipe, this may be more than offset by savings in the
installation cost. Work boats or special-purpose vessels
equipped with large-diameter
reels can lay long lengths
of flexible pipe in short periods of time. Flexible pipe normally does not require a separate coating on the outside.
but it may require cathodic protection of the end connections.
All flowlines should be protected from abrasion and
physical damage from other crossing pipelines, and expansion loops may be necessary if the installed configuration does not allow for expansion and contraction from
temperature changes. Other considerations include possible scour damage or vibration fatigue where bottom currents exist and high stresses where the line bridges low
spots on uneven seafloor. Installation
methods for
flowlines are generally the same as for other subsea pipelines. These are discussed in the Offshore Pipelines
section.
Control lines, both hydraulic and electrical, for subsea
well completions
are discussed in the Electrical, Instrumentation and Control Systems section.
Well Servicing-Wireline
vs. Through Flowline. There
are two common techniques for performing downhole
work on subsea wells when the work tasks do not require
removing the Christmas tree and tubing. The most common procedure is to install a workover riser between the
surface vessel (drilling or workover rig) and the top of
the Christmas tree above the swab valves and to install
a wireline lubricator on top of the riser. Conventional
wireline tools and equipment can be used to remove and
to install safety valves and gas-lift valves, to shift sliding
sleeves, or to make temperature and pressure surveys.
Vessel heave must be compensated for, but otherwise the
procedure is the same as for land and platform wells.
To overcome the delays and high cost associated with
wireline work on subsea wells (vessel availability, high
daily rates, and weather delays), a procedure for servicing wells remotely from the process platform has been
developed. It is called through flowline (TFL). It is
basically a set of tools that are inserted into the flowline
at the platform and hydraulically
pumped through the
flowline, through 5-ft-radius flowline loops (bends) at the
subsea Christmas tree, and down the tubing. A complete
hydraulic loop is required between the platform and the
well to pump the tools down and back. This means that
a second flowline is necessary for work to be performed
at or immediately below the wellhead and a second tubing string is required, with controllable communication
between the tubing strings downhole to perform work
downhole. TFL tools are available to fit common tubing
sizes and to perform virtually all the same tasks as wireline tools. Numerous technical papers have been written
on TFL tools and techniques.
Gas wells generally are not suitable candidates for TFL
servicing because a gas-free hydraulic loop is required

PETROLEUM

ENGINEERING

HANDBOOK

for circulating
the tools. A decision favoring either
wireline- or TFL-servicing
procedures for an oil well
should be based on a full evaluation of operating conditions, anticipated downhole service requirements,
the
availability of trained personnel, and an economic comparison of installed costs and servicing costs.
Well Workovers. Work on subsea wells that requires the
tubing to be pulled or is otherwise beyond the scope of
wireline or TFL tools is a major undertaking and requires
extensive planning and preparation.
Unless a specialpurpose vessel is available that is suitable for workovers,
a semisubmersible
or ship-shaped drilling rig must be
scheduled and mobilized with a workover riser and running tools specially designed for the Christmas tree and
tubing hanger. This equipment is needed in addition to
a regular drilling riser and BOP stack, which must have
a hydraulic connector that is compatible with the wellhead. Space limitations on some drilling vessels may
preclude their use, which further complicates
the
workover.
Floating Production

Facilities

For many years, floating drilling rigs (semisubmersibles,


ships, and barges) have conducted drillstem tests and
short-term production tests of newly drilled wells. These
wells were nearly always drilled for exploratory or delineation purposes. Common practice was to abandon the
wells temporarily or permanently after testing. Following the installation of a platform, development wells were
then drilled and produced. Because brief production tests
frequently provided insufficient reservoir data, and because delaying well production until platforms could be
fabricated and installed resulted in poor cash flow, the
concept of floating production facilities (FPFs) was developed. This concept requires some type of floating vessel (ship, barge, or semisubmersible)
that is equipped for
processing crude-oil production instead of for drilling. The
vessel is either moored in place with multiple anchors,
or is connected to some type of single-point mooring
(SPM) or articulated tower. Crude-oil production from
one or more seafloor wells is produced to the FPF either
directly through individual
pipe risers or through a
seafloor manifold center and multiple-bore riser assembly (Figs. 18.37 through 18.39).
Applications.
Two applications for FPFs were mentioned above: long-term production tests and accelerated
(early) production. The use for long-term testing nearly
always involves only one well and may be for a duration
of anywhere from 2 to 3 months to a year or longer. Some
reservoirs, particularly limestone as opposed to sandstone,
cannot be evaluated from a short-term test, and full field
development may be unacceptably risky without extended test data. The simplest FPF installation for production
tests is a vessel moored with multiple anchors directly over
a seafloor well connected by drillpipe or tubing to a subsea test tree installed in a BOP on the wellhead. A similar installation generally preferred for longer-term testing
would use a flexible pipe riser instead of drillpipe or tubing and a conventional subsea Christmas tree instead of
a BOP with test tree.
The same installations described above can also be used
for early production if only one or two wells are involved.
For more wells (up to 5 or lo), a seafloor manifold center

OFFSHORE

18-35

OPERATIONS

Fig. 18.37-FPF

with flexible

pipe risers and floating

may be used with a multiple-bore riser and some type of


SPM or articulated tower. An SPM or tower has the added
advantage of permitting the process vessel to weather-vane
into the wind and seas and thus to mitigate the effect of
bad weather on process equipment.
A third application for FPFs is permanent process facilities for a small or short-lived oil field where a conventional platform is either uneconomical or marginally

Fig. 18.38-FPF

with seafloor manifold,

composite

loading

hose for two subsea wells.

economical. Particularly where reuse of the facility may


be a factor, the ease of moving an FPF to a different location or field may have a significant impact on overall
economics.
Semisubmersibles
vs. Tankers. Most FPF installations
have been converted semisubmersible drilling rigs or convetted oil tankers. Semisubmersibles are characteristically

riser, and seafloor pipeline to loading buoy and tanker.

PETROLEUM

18-36

Fig. 18.39-Flow

diagram

for FPF process equipment

better suited to severe weather areas such as the North


Sea, because of their superior stability. Their main drawback from an operating standpoint is the lack of sufficient
oil storage capacity to prevent shutdowns when tankerloading operations are disrupted. Loading disruptions are
not uncommon and can result from either equipment
failure or bad weather.
In relatively mild weather areas, converted tankers perform satisfactorily as FPFs, especially when moored to
SPMs or loading towers that permit the tanker to weathervane. Large oil-storage capacity, inherent to tankers,
greatly facilitates the scheduling of shuttle tankers to transport the crude oil to refineries or terminals.
The choice between semisubmersibles
and tankers for
FPF service is heavily influenced by the availability of
surplus vessels, the operating conditions, and the geographic area. A careful investigation of the used-vessel
market is essential to an economic decision. At the time
of this writing, new special-purpose vessels designed specifically for FPF service are being promoted. These include semisubmersibles
with some storage capacity to
offset that drawback and ship-shaped vessels with antiroll devices to improve stability. A wide selection of FPF
designs and vessels probably will make their use more
economically and operationally attractive in coming years.
Disposal of Oil, Gas, and Water. Technically, there is
no reason why oil cannot be shipped from an FPF by pipeline. Most installations to date, however, have transported oil with shuttle tankers. Loading operations can be
accomplished by a floating hose directly between the FPF
and the shuttle tanker. This inexpensive approach is practical in mild weather areas, especially with low production rates. For higher production and transfer rates and
for adverse weather conditions, a seafloor pipeline or hose
between the FPF and a dedicated loading buoy is safer
and generally will result in less downtime.
Water produced with crude oil can be treated and disposed of the same as on a platform. It can be cleaned and

for one-well,

long-term

ENGINEERING

HANDBOOK

test.

treated for reinjection into the reservoir, or, where local


regulations permit, it can be cleaned and discharged into
the ocean. Depending on downstream terminal or refinery
conditions, shipment of small amounts of water production with the crude oil may be possible.
Small amounts of gas production have to be flared in
a typical FPF installation. If flaring is not permitted, or
if economics favors reinjection, compressors can be installed for this purpose. Use of FPFs in fields producing
large amounts of gas may require additional facilities for
gas treating, processing, and disposal.
Offshore

Pipelines

The design and installation of subsea pipelines bear only


slight resemblance to their counterpart activities on land.
Preliminary
sizing of lines can be based on generalpurpose pressure-drop curves as long as the effects of the
ocean environment
on fluid rheology are understood.
Also, preliminary cost estimates can be made on the basis of either estimating manuals prepared for this purpose
or historical data for similar installations. Final pipeline
designs, detailed plans, and cost estimates for fabrication
and installation, however, are best handled by pipeline
contractors or consultants who specialize in this activity.
Flowlines. Flowlines for subsea wells range in size from
2 to 6 in. As indicated previously in the Subsea Completions section, conventional steel pipe is used for most installations. It is readily available and does a good job when
protected against corrosion and physical hazards. Flexible pipe made of laminations of steel wires and other materials is now available from several sources and has been
used in a number of instances. It is made in a range of
pressure ratings and a variety of materials that are suitable for most applications.
Welded flowlines sometimes are made up on the beach
and then towed to the point of placement. Towing can
be at or near the surface of the water with the pipe supported by buoys or other buoyant material, or it can be
just off-bottom by a combination of buoys and chains for

OFFSHORE

OPERATIONS

18-37

SIDE VIEW

IWZ-~ODAT~ON

FRONT VIEW

Fig. 18.40-Conventional

buoyancy control. More common methods of installation,


however, are either by conventional
lay barges (Fig.
18.40) or by reel barges (Fig. 18.41). The former make
up straight lengths of pipe on the barge and feed it into
the water by way of a curved stinger as the barge is
winched along the flowline route. The purpose of the
stinger is to control radius of curvature as the pipe is lowered into the water and thus to prevent buckling. Buckling and overtensioning
of the pipe as it contacts the
seafloor are prevented by maintaining a predetermined
amount of tension on the pipe as it leaves the barge and
by controlling the forward movement of the barge.
Probably
the most popular method of installing
flowlines, both conventional steel pipe and flexible pipe,
is with special-purpose
reel ships or reels mounted on
large work boats. Depending on pipe diameter, several
miles of pipe can be reeled onto one or more large reels
at a shore-mobilization
site and then rapidly reeled off
at the placement site. The main advantage of this t&hnique is the speed of installation. Fast installation reduces
not only the number of offshore construction days but also
costly interruptions caused by bad weather. A job that
might require a week of offshore construction time with
a conventional
lay barge is much more susce tible to
weather downtime than a job that can be camp Peted in 1
or 2 days with a reel ship.
Flowline connections at platforms generally are made
by pulling the line up through a curved conductor pipe
called a J tube and then securing the line at the platform
deck opening with a flange or clamp. Several procedures
are used for connections to subsea wells, depending on
Christmas tree configuration and whether the flowline in-

DFCK

LAWUT

pipe-lay barge.

stallation starts at the tree (a first-end connection) or ends


at the tree (a second-end connection). Most tree connections have been a pull-in type, where the flowline is first
laid on the seafloor and then pulled into a receptacle on
the tree base with a wire rope. One advantage of this is
that it can be performed either on a first-end or secondend connection. A pull-in procedure can also be used as
the flowline is being lowered to the seafloor, making a
connection at the tree base before laying the flowline on

Reel Capacity

Fig. 18.41-Work

boat with pipe-laying

reels

PETROLEUM

18-38

ENGINEERING

HANDBOOK

-_r

RUN FLOW LINE POSITIONER

LAND FLOW LINE POSITIONER

REMOVE GUIDE LINES,;;;:1


STINGER. BEGIN LAYING FLOW LINES

Fig. 18.42-J-lay

method for installing

RETRIEVE FLOW LINE


POSITIONER RUNNING TOOL

ACTUATE FLOW LINE CONNECTOR.

flowlines

the seafloor. A procedure particularly applicable to subsea wells in very deep water is the J-lay first-end connection (Fig. 18.42), where the flowline is run vertically to
the tree base from the drilling rig, stabbed into a receptacle, and then laid down into a horizontal position as the
drilling rig or pipe-lay vessel moves away from the wellsite toward the platform. A trunnion-type assembly on the
end of the flowline permits the line to be laid down without
bending. The mating of the flowline to the Christmas tree
is made after the line is fully horizontal.
Many different devices and procedures have been developed for making the actual connection between the
Christmas tree and the flowline. In shallow water, divers
can install a piping spool with flanges or couplings on each
end. Diverless connections usually have some type of
hydraulically actuated device that is operated remotely
from the drilling rig.
Larger Pipelines. Pipeline diameters from 8 to 36 in. or
more are used extensively for the transfer of crude oil
and gas from offshore fields to land sites. Installations
typically have been made with conventional
pipe-lay
barges as described previously, but reel barges also are
used extensively for the smaller sizes. Both bottom tows
and surface tows are used in limited applications where
logistics favor them. As discussed previously, a pipeline
can be the single most expensive element of an offshore
installation, sometimes exceeding the cost of one or more
platforms. Numerous factors must be considered when

away from subsea well (first-end connection).

designing a line and planning its construction to minimize


installation difficulties and to ensure satisfactory operation throughout the expected life of the line. Many technical papers and magazine articles have been written about
this subject and are excellent
sources of further
information-e.g.,
proceedings from the annual Offshore
Technology Conference.

Arctic
Production operations in the offshore Arctic regions are
within the reach of existing technology. Procedures used
onshore and offshore in less hostile regions, however,
must be modified to meet the challenges of the harsh climatic conditions in these remote locations. I5
In the last decade, the major areas of industry interest
have been the offshore regions of Alaska I6 and Canada.
The environmental
conditions vary significantly in each
of these regions. Major factors that affect normal offshore
operations are extremely cold temperatures,
fog, gusty
winds, short open-water seasons, permafrost, and the persistence of ice. The specific production system that is
selected must be tailored to each unique combination of
these factors to ensure safe oilfield development.
Environmental

Conditions

Ice Characteristics. Sea ice is the principal environmental


factor in all of the offshore Arctic areas. I7 The most
abundant type of sea ice that is encountered offshore is
less than 1 year old. This first-year ice begins to form

OFFSHORE

OPERATIONS

during fall and grows to a thickness of 4 to 8 ft during


the winter. Sheets of ice close to shore become landfast
and remain locked in place throughout the winter. Beyond
the landfast zone, the ice is kept in constant motion by
wind, currents, and, in some areas, the influence of the
Arctic polar pack. This dynamic movement causes shearing and impacting between ice features that produce ridges
of ice several miles in length. Ice ridges formed in this
manner are called pressure ridges. Localized ridging
around a grounded ice feature, the shoreline, or a structure is considered a rubble pile. In areas of extremely cold
winter temperatures, the ice blocks within a ridge or rubble pile begin to refreeze into a contiguous feature. Depending on the conditions,
the refrozen consolidated
thickness could become several times larger than the firstyear ice thickness.
Sea ice that survives one or more melt seasons is considered multiyear ice. The predominant source of multiyear ice features is the polar pack. The pack consists
primarily of floes 2.5 to 50 ft thick with embedded ridges
50 to 100 ft deep. During the summer, northerly winds
break off portions of the pack and push them toward shore.
These multiyear floes are commonly 1,000 to 2,000 ft in
diameter.
The other major type of ice is not formed from seawater but is freshwater ice from the glaciers of Northern
Canada. In the Arctic Ocean, the glacial fragments are
called ice islands. These tabular-shaped features are several thousand feet in diameter and more than 200 ft deep.
Because of the enormous size and slow movement rates
of these features, they can be tracked for several months
before encroachment upon a given area. Most of the areas
of interest for oilfield development in the Arctic Ocean
are also in relatively shallow water. This shallow bathymetry causes these freshwater ice features to run aground
before they are a threat. In the North Atlantic, similar
glacial features are called icebergs. Icebergs that weigh
more than 50 million tons have been observed in water
depths beyond 1,600 ft. Again, the local bathymetry dictates the maximum size iceberg that could encroach upon
a given area.

Ice Loading. Ice exerts the predominant forces on Arctic offshore structures. Extensive laboratory and field tests
have been conducted on small- and large-scale specimens
to determine in-situ strength characteristics for design.
From the results of these tests, the mechanical properties
of ice are predicted that consider its salinity, temperature,
crystallographic
structure, and loading rate.
Newly formed sea ice is relatively warm-only
a few
degrees above seawater temperature-and
high in salinity. As the ice sheet grows, the temperature at the surface
reduces to the ambient air temperature, while the bottom
remains near the seawater temperature. The salt in the
crystal structure of the sheet begins to consolidate into
brine droplets. These droplets migrate down through the
thickness of the sheet, creating drainage channels and lowering the overall salinity. Fresh water, from precipitation
or melting snow cover, fills these channels and refreezes,
thereby further reducing the salinity. The result is an ice
feature with varying strength, strongest at the surface and
reducing through the thickness with increasing salinity and
warmer temperature. For multiyear ice, this process ap-

18-39

proaches equilibrium. The multiyear ice floe is near zero


salinity and has a relatively cold temperature through its
thickness. This results in an ice strength several times
greater than a first-year ice sheet.
Another parameter that influences the strength of sea
ice is the loading rate. When ice is loaded at a very slow
strain rate, it exhibits a plastic behavior. Loaded rapidly,
it behaves as a brittle material. Empirical equations have
been developed that relate the ice movement rate and
shape of the structure or indenter to the strain in the ice
feature.
The shape of the structure is also a primary factor in
producing a crushing, buckling, or flexural failure of the
ice feature. For narrow structures relative to the ice thickness, crushing is the predominant failure mode. As the
width increases, a combination of crushing and buckling
of the ice field around the platform results in the development of a rubble pile. This rubble pile will then shield
the structure from direct impact of subsequent ice floes
and ensure failure of the ice mantle away from the production facility. And finally, sloping-sided structures normally force a flexural ice failure. Because ice flexural
strength is 20 to 40% of the crushing strength, an appreciable reduction in ice forces can be achieved when
a bending failure is induced.
Waves. The wave conditions in the Arctic are similar to
other offshore areas, and the design of structures or islands against wave loading is well established. Nearshore
sea states can be defined by determination of the openwater area along the storm route or fetch and the water
depth. In the Arctic Ocean, the presence of sea ice and
the polar pack limits the open-water fetch for storms to
generate and consequently reduces the design wave height.
Breaking wave conditions exist in most shallow-waterdepth locations and around gravel islands. This diffraction, shoaling, and refraction of the waves produces highly
irregular sea states. Because the interaction of the waves
and structure is dependent on structural geometry, the
forces the design wave exerts on the structure should be
determined by model testing or approximated by linear
diffraction analysis.
Scour of the foundation around the structure caused by
waves must also be considered and appropriate protection methods implemented.
The scour protection could
consist of rock, sand bags, or precast concrete elements
placed around the structure after installation. For artificial islands, this becomes slope protection and can be used
to reduce wave and ice run-up by providing an artificially rough surface.
Permafrost. Permafrost is soil at a temperature below
32F with partially or completely frozen pore water. Drilling and producing operations in areas with permafrost
have been well defined from the experience of the Prudhoe Bay field. In most nearshore areas of the Arctic, permafrost has been found at or near the mudline. These soils
are very stiff and can make excavation for pipelines or
driving of piling nearly impossible. Permafrost normally
is soil bonded by ice and is very susceptible to changes
in temperature. This can result in significant changes in
the soil characteristics
and must be considered in the
design.

PETROLEUM ENGINEERING HANDBOOK

18-40

Fig. 18.43A Protected-slope production island.

Production Structures 16
Artificial Islands. Artificial islands already are used in
many shallow-water areas throughout the world for permanent drilling and producing facilities. The islands that
are currently being used for drilling in the Arctic consist
of either unretained or retained beach slope systems, as
shown in Figs. 18.43A and 18.43B. Because of the short
summer construction season and, in some areas, the lack

of island fill material, the quantity of fill required for the


island should be minimized. The minimum island working surface is determined by the area required for drilling and production operations. To reduce the quantity of
island fill, the steepest side slopes that the mode of construction and fill material will allow should be provided.
The minimum side slopes of unretained islands depend
on whether the island is constructed by summer dredg-

Fig. 18.43B Caisson-retained production island

OFFSHORE

18-41

OPERATIONS

+ermafrost

cement

Fermafrost

cement

cement

Permafrost

Non
ydraulic

set

liner

cement

freezing

fluid

hangers
Safety

valve

Non freezing
insulating
fluid
-Class

G cement
Stage

KeStage

collar

collars
Permafrost

cement

hanger

-+ 9000
TV0

Class
--Class

---Safety

G cement

G cement
joint

Hydraulic

set

packer

i 9700'
TVD
BP Alaska

hole-casing

Arco

program

Fig. 18.44-Hole

casing programs,

ing or winter transport of onshore borrow material over


the ice to the desired location. The side slopes for summer dredging are approximately
1 : 20 (vertical to
horizontal), and for winter construction are 1 : 3. On completion, sandbags or concrete mats are placed on the exposed slopes of the island to prevent ice and wave erosion.
Sandbags, stiffer soils for embankments, or caisson units
are used on retained islands during construction to reduce
the required volume of fill. The caisson units typically
consist of vertical walled concrete or steel units. The caissons also provide easy access to the island as a dock for
resupply and could be used for storage of consumables
or oil.
Artificial islands must be designed to withstand the
horizontal forces exerted by ice. The potential failure
modes of the island consist of slope instability, bearing
failure, or horizontal shearing of the island near the waterline. Each of these failure modes can be predicted by classic geotechnical analysis. The only variable in the analysis
is the properties of the island fill material. During winter
construction, the fill is delivered to the site at the cold
ambient temperature and dumped into the sea. Ice forms
on the granular material and inhibits consolidations.
As
the island surface thaws, considerable settlement may take
place. To minimize the effects of thaw settlements, thermal analysis of the freezing and thawing interface should
be conducted to determine the proper gradations of fill
material.
The design of production facilities placed on an island
is similar to that on land. Equipment foundations must
be designed and insulated to reduce the potential for frost
heaving, pile jacking, and thaw settlements from seasonal
thawing and freezing of the island surface and, in some

hole-casing

program

Alaska.

areas, subsea permafrost. To prevent thaw settlement, an


artificial refrigeration system for the fill material could
be installed. Placement of equipment and accommodation
modules should account for predominant wind, ice movement, and wave directions to ensure safe year-round operations.
The well systems should be vertically drilled through
the frost-susceptible
island surface and permafrost and
then directionally drilled to true vertical depth (TVD).
Wells should be spaced close together to minimize the
overall size of the island surface and to reduce the effects
of thaw subsidence. The casing program should be designed to withstand freeze-back loading during periods
of well inactivity and to accommodate differential movement in the tubing string owing to the thermal effects of
the drilling fluids. Two casing programs in the Prudhoe
Bay field are shown in Fig. 18.44, I8 which shows that
both systems incorporate the use of a permafrost cement
and provide a safety valve in the production tubing string
below the permafrost for emergency shut-in.
Gravity Structures. Various types of gravity structures
are being proposed for use in the Arctic. Many of the conventional gravity structures that are used in the North Sea
are being adapted for the deepwater and moderate-iceconcentration areas. In the more hostile areas of the high
Arctic, vertical- and sloping-sided gravity structures are
being proposed. These structures provide the large deck
load and space requirements, protection of the wells within
tower shafts, and storage of oil. Because of the extreme
winter ice conditions in many areas, the production facilities will have to operate 9 months without major
resupply.

18-42

PETROLEUM

Fig. 18.45-Vertical-sided

The vertical-sided structures (Fig. 18.45) are proposed


for the shallow, nearshore areas in the Arctic. These structures typically are rectangular or hexagonal and are capable of being installed directly on the seabed or subsea
berm. Production equipment can be placed directly on the
working surface of the top slab or integrated into the hull
of the structure. Wells are drilled and produced directly
from the deck of the structure. Because of the large width
of this concept, the structural integrity of the system is

ENGINEERING

HANDBOOK

structure.

not sensitive to local discontinuities


in the seabed from
ice gouges or settlements in the foundation from local
degradation of permafrost.
Conical, sloping-sided structures (Fig. 18.46) are being
proposed for the deeper-water,
dynamic-ice-movement
areas. This geometry induces flexural failure of the ice
features and is relatively transparent to pack-ice movements. The deck is fully outfitted with processing equipment before it is mated with the structure. Wells are
confined to a central moon-pool area in the cylindrical
throat. Consumables and oil can be stored in the base.
Piled Structures. Piled steel structures have been developed primarily for the Bering Sea area of offshore Alaska. These structures are similar to conventional template
or jacket concepts but must be modified to resist annual
sheet-ice loading. A typical geometry is shown in Fig.
18.47. The platform concept consists of four or eight main
pile legs with intermediate bracing of the legs omitted in
the ice-loading zone near the waterline. Well conductors
and oil-transport lines are positioned within the legs of
the platform for protection from ice loading. This requires
close spacing of the wells and, in some cases, completion of the wells at different levels of the deck. Diveraccess tubes may also be located in the legs to facilitate
the repair and inspection of subsea components of the platform during complete ice coverage.
In most other Arctic areas, pile structures are not practical. Subsea permafrost makes pile installation nearly impossible. The short construction
season also does not
accommodate the installation, pile driving, and placement
of the topside modules in one season. Also, the hookup
and commissioning of the production equipment modules
would be very expensive in these remote areas.
Transportation

Fig. 18.46-Arctic
mobile
structure.

drilling

structure:

sloping-sided

Systems

Pipelines. Offshore pipelining is the predominant mode


of crude-oil transport proposed for Arctic regions. The
pipelines will interconnect platform facilities, mooring

OFFSHORE

18-43

OPERATIONS

structures, and land-based facilities in much the same manner as in conventional offshore locations. The principal
factors affecting the construction and operations of Arctic pipelines are the short open-water seasons, subsea permafrost, scour, and ice gouging of the seafloor.
Pipelines can be placed directly on the seabed in deep
water, in trenches in areas susceptible to ice gouging or
scour, and on causeways or elevated bridges at shore
crossings. In areas of extreme ice gouging, redundant lines
may be used to lower the risk of interrupted production.
In deepwater, moderate-ice areas of the Bering Sea, the
lines will be laid directly on the seabed by conventional
lay-barge methods. The 6- to 9-month open-water season, extreme summer wave conditions, and logistics of
operating several hundred miles offshore reduce the efficiency of the construction process. The distance from
shore or an offshore loading terminal may also require
pump stations along the pipelining route.
Marine pipelines in the Arctic Ocean are considered
feasible but will require the greatest challenge to existing
technology. The open-water season lasts approximately
1 to 3 months a year. In shallow-water locations, the keels
of ice features may gouge the seafloor for several miles
in a single ice-movement event. Pipelines must be buried
in trenches that are deep enough to ensure no damage.
Offshore permafrost may also cause difficulty in obtaining the desired trench depth and may require the use of
cutter-suction dredges to remove the ice-rich soil. Once
the pipeline is installed, refrigeration of the trench may
be required to ensure no subsequent thawing of the permafrost. Nearshore pipelines may also have to be designed
for wave-induced erosion or strudel scour. Strudel scour,
which is common at the mouth of rivers, is the process
in which river-water outwash flows over the nearshore
sheet ice and floods down through holes and cracks in
the ice. This jet of water could create large, eroded pockets in the seabed and produce long, unsupported spans
of pipe.
Marine Terminals and Tankers. Transportation
of
crude-oil products from many of the remote Arctic oil
fields also can be accomplished by offshore terminal and
tanker systems. The systems could consist of an interfield
pipeline-gathering
network with processed crude shipped
directly to shore. Once onshore, the crude could be stored
until it is loaded onto tankers and shipped to market.
Another form of marine terminal system is direct shipment of processed crude from offshore platforms to
tankers through single-point mooring systems.
The design, construction,
and operation of these systems have been proved in many sub-Arctic areas. However, some of the components must be modified for the
cold temperatures and persistence of ice. Loading arms
must be designed to ensure that hoses remain elevated
above the ice mantle and can accommodate loading of
tankers from any location around the terminal. Tankermooring systems must be designed for both wave- and
ice-loading conditions. Extreme ice areas may require offshore loading directly from fixed platforms or from subsea facilities, making maneuvering
and stationkeeping
very difficult. Icebreaker assistance vessels may also be
required to ensure safe access and departure from the terminal by tankers.
To travel through the offshore Arctic regions, purposebuilt icebreaking tankers may be required. The amount

Fig. 18.47-Piled

structure.

of hull stiffening will be dictated by the ice conditions for


its area of operation. The modifications could range from
simply thickening the hull plate to the requirement of an
icebreaking bow and turbine-powered propulsion system.
Internal bulkhead arrangements should also be arranged
to ensure that oil storage tanks are adequately stiffened
and not susceptible to direct ice impact.
Special Considerations
Ice Management. A critical support system, unique to
the Arctic, is the ice-management
system. The key component of this system is instrumentation
of the structure,
foundation, and ice field around the structure to monitor
both local and global ice loading. Other elements of ice
management include icebreaker support vessels for tankers
or supply boats, tractor-mounted
ditch diggers for slotting the ice to reduce loads, and water-pumping units to
flood areas around a platform to stabilize ice movement.

Electrical, Instrumentation
Control Systems

and

Offshore production facilities have the same basic requirements for electric power and control systems as onshore
facilities. These are a power source with a reliable distribution system, instrumentation
to control and to monitor
production operations, and a safety shutdown system
tailored for the installation.
Although offshore and onshore installations have similar needs in these areas, the two operations are significantly different in other respects. As pointed out earlier,
deck size and payload significantly affect offshore structure costs. Consequently, offshore electrical facilities must

PETROLEUM

10-44

be designed for minimum weight and space while still


offering a high degree of flexibility, reliability, access,
and maintainability.
Because of deck space and layout
limitations, hazardous area considerations are more complex and often are governed by different regulatory codes
depending on the type of structure, its location, and the
responsible governmental agency. Additionally, environmental considerations are much more demanding because
of the generally salt-laden atmosphere and the possibility
of saltwater washdowns and sea spray.
Several wiring methods are approved for offshore installations, but the method used frequently depends on local practices and personal preferences. There is no single
correct method for wiring an offshore installation as long
as the appropriate codes are satisfied and the craftsmanship is of average quality or better.
Offshore-production
instrumentation
and control systems also tend to differ from their onshore counterparts.
Offshore oilfield operations are confined to a relatively
small spot on the ocean rather than being spread over
several acres. Consequently,
controls tend to be much
more centralized. This trend toward centralization has become more pronounced with increased use of computerbased production-monitoring
and control systems.
Safety shutdown systems onshore and offshore are
roughly equivalent. However, offshore requirements regarding functionality and reliability are somewhat more
stringent because of the greater potential for spills and
resultant pollution. Increased concern for personal safety and environmental
protection creates a more stringent
atmosphere for approval, testing, and inspection by outside agencies, as well as thorough documentation of the
entire process and facility by the operator.
Alternative methods of addressing and satisfying these
and other considerations peculiar to offshore operations
will be discussed. The intent is to lay out the problems,
to highlight areas of concern, and to discuss possible solutions, rather than to present specific, detailed engineering methods for design and installation of electrical and/or
control systems.
Codes and Regulatory

Authorities

Offshore electrical installations are governed by one or


more codes and regulations, depending on the type of offshore structure involved and its location. In U.S. waters,
electrical designs are governed by local or state electrical codes and in most cases the Natl. Electrical Code. I9
Generally, the installation should be in accord with the
more stringent requirements of the applicable codes. The
guidelines given in API RP 14F* provide direction for
accepted good practice in accordance with most applicable codes. If the facility falls under U.S. Coast Guard
jurisdiction (e.g., TLPs, mobile offshore drilling units
and FPFs), all or portions of the electrical system also
must be designed, installed, and operated in accordance
with the applicable portions of U.S. Coast Guard Regulations 46 CFR Chapter I, Subchapter IA, Mobile Offshore Drilling
Units,
Part 108 and Subchapter
J-Electrical
Engineering, Parts 110-l 13. *t Overseas installations frequently fall within the jurisdiction
of local/national codes or regulatory agencies. This situation
occurs in the North Sea where Lloyds of London or Det
norske Veritas frequently are named as certifying authority, and their requirements must be met. If there is no lo-

ENGINEERING

HANDBOOK

cal certifying authority, it generally is good practice to


design all systems to meet normal U.S. requirements for
that type of facility.
Platform

Loads

With the exception of the hotel loads that serve accommodations and personnel needs, normal platform loads
are not greatly different from the normal assortment of
onshore loads. Processes are likely to be concentrated in
a single location offshore, and the loads are appropriately
higher. Loads usually consist of a mix of transfer pumps,
compressors, fans, and heaters as well as utilities (air compressors, sump pumps, fire water pumps, water makers,
sewage facilities, etc.). Other typical loads include pumps
and gas compressors associated with shipping, artificial
lift, secondary recovery, and pressure-maintenance
operations. Depending on platform type and production facilities and rates, the loads on any given platform in a
field can range from less than 25 kVA to more than 40
MVA.
Layout of Facilities
Considerations that govern layout of offshore electrical
facilities are similar to those for land installations, but the
options are more restricted because of the space limitations. The primary requisite is to separate to the maximum extent possible the sources of ignition from the
process facilities. Electrical equipment should be kept out
of hazardous areas when economically feasible. Primary
electrical switch gear is frequently grouped in a pressurized, central electrical-equipment
room. Remote motorcontrol centers (MCCs) sometimes are used to reduce
the amount of platform wiring. Installing MCCs near load
concentrations
and supplying the MCCs with highvoltage feeders shortens branch circuits that feed individual loads. Remote MCCs frequently must be purged or
pressurized because of their location. The design approach
appropriate in each individual case must be based on practicality and economics.
Detailed load analyses must be made early in the design stage. Each analysis must consider both initial and
anticipated future load. Results of the analysis should control design of primary power facilities, layouts of conduit or cable ways, switchgear space, and distributionequipment layouts. Designs must allow for expansion in
each of these areas. The designer must carefully consider
possible future system expansion and pay particular attention to possible future artificial-lift and water-injection
requirements. Hydraulic pumping and electric submersibles can add significantly to the ultimate electrical load.
It is important to consider the potential for future changes
in facilities as reservoir conditions change and possible
increased demand on the electric power system.
Primary

Electric

Power

Typically, offshore facilities either generate power locally


or they are fed by submarine cables. Generated power
ranges from 480 to 4,160 V. The higher voltage levels
are more prevalent where there are high horsepower loads
or where platform drilling rigs are not set up to generate
their own power. Very large platforms are sometimes designed with primary power system voltages of up to 13.8
kV.

OFFSHORE

i a-45

OPERATIONS

The primary source of generated power usually consists of one or more brushless synchronous generators with
either diesel-engine or turbine prime movers. Turbines
may be gas- or diesel-fueled or they may be dual fueled.
Dual-fueled turbines generally are set up to run on diesel
initially and then to cut over to produced gas as production builds to a stable supply. Turbines offer the advantages of lighter weight and the opportunity for the use of
waste heat in oil production processing. Turbine-generator
packages, however, are very costly, require more maintenance, and in some cases, may require more deck space
than internal-combustion-engine-driven
generators of the
same capacity. In addition, primarily because of the somewhat limited selection of turbine sizes available, it can
be more difficult to match turbine generator packages to
the electrical system loads.
The second source of offshore power, submarine cable, offers an efficient means of electrification if adequate
sources are located nearby. Depending on the loads and
the distances involved, cable-system voltages can range
between 480 and 35,000 V. Submarine-power-cable
technology is well established. Power cables generally include
one or more pairs of small-gauge wires dedicated to
telecommunications
or telemetry and also are protected
with torque-balanced
double layers of galvanized armor
wires that may be plastic-coated for additional protection.
Power cables may be buried, depending on local conditions or regulations. Tubular risers are required at each
platform served by submarine cables to protect the cables as they rise from the seabed to the platform deck.
The riser generally should be filled with a corrosioninhibiting fluid to preserve the long-term integrity of the
armor wires. When the submarine cables originate on
land, they must be buried and suitably protected through
the surf zone. Mechanical protection is frequently provided by additional armor or by the installation of heavywall pipe through this area.
Secondary/Back-up

Power

Essential loads on offshore facilities must remain energized even when the main power source fails. These loads
usually fall in the category of navigation lights (ship and
aircraft), foghorns, communications,
emergency lighting,
and possibly some selected hotel loads. In some instances
where marine regulatory bodies have jurisdiction,
the
emergency system may have to be expanded to serve other
loads related to the safety of personnel on board.
High-volume oil and gas facilities are frequently designed to shut in completely in the event of power failure,
in which case process or shipping loads do not have to
be considered in determining backup power loads. Some
smaller facilities, such as wellhead platforms, are designed
to continue production operations without any electrification other than lighting, navigation lights, and foghorns.
Emergency power generally is derived from onboard
generators, storage batteries, or both. The generators
usually are small (less than 500 kW) and diesel-engine
driven. Diesel engines that are used in emergency service normally are equipped with automatic starters. These
engines also are equipped with cooling water and lubeoil heaters to ensure that they will start reliably and instantaneously
(within 10 seconds) on loss of primary
power.

Fig. 18.48-Typical
electrical one-line
emergency generator.

diagram

with ups and

Offshore facilities with electronic instrumentation


or
computer-based monitoring or telemetry systems generally need uninterruptable
power supplies (UPSs) to provide reliable, clean power for these loads. A UPS is
essential on platforms where some loads are supplied from
onboard, solid-state, silicon-controlled
rectifiers (SCRs)
because of the high level of electrical noise injected into
the power system by the SCRs.
The emergency bus typically is tied into the normal
power bus through an automatic transfer switch. The
emergency bus receives power during normal operation
from the main bus. On failure of the primary power, the
transfer switch opens the tie between the two buses to isolate the emergency bus and its loads from the remainder
of the platform loads. Additional controls automatically
start the emergency generator to energize the emergency
bus. Normally, no provisions are made for operating the
emergency generator in parallel with the main power supply. Interlock circuits should be provided, however, to
permit testing the emergency generator offline without
energizing the emergency bus.
Typical one-line diagrams for a UPS and an emergency generator bus tie are shown in Fig. 18.48.
Distribution

System

Offshore power-distribution systems and associated equipment do not differ substantially from land-based operations. Depending on load sizes, distribution voltages will
normally range from 120 to 2,400 V. Distribution systems with 4,160 V and higher are rare except on very

PETROLEUM

18-46

large platforms with high horsepower loads. As with landbased systems, accepted good practice and system capacity
determine the maximum allowable horsepower for individual loads and the appropriate supply voltages.
Motor-control centers and switchgear usually are installed inside pressurized or purged enclosures or modules. The pressurizing or purging frequently is required
to ensure a nonhazardous operating environment for the
electrical equipment so that air-break switching equipment
can be used safely. Code requirements for purging are
prescribed in the Natl. Electrical Code I9 and Natl. Fire
Protection Assn. Bull. NFPA 496. * As indicated in the
references, interlocks usually are provided either to sound
an alarm or to shut down all supplies when a loss of pressurization or purge occurs. Depending on the particular
operations involved, a simple alarm may be permissible
if shutdown of the process could create a greater hazard
than continuing operation during a short-term loss of pressurization or purge protection. Each installation is sitespecific and must be considered on its own merits.
System transformers may be installed either indoors or
outdoors. Both single- and double-ended line-ups are used
offshore, depending on the nature of the operation.
Double-ended 100%~capacity supplies with bus tie breakers are frequently used for increased reliability and continuity of service in the event of equipment failures.
Hazardous

Areas

The possibility of fire or explosion because of the ignition of leaking gas or liquids is a concern in all producing facilities. The concern is even greater offshore because
of the concentration of personnel and facilities in a relatively confined area where fire can be difficult to extinguish and platform evacuation can be a complex problem.
The requirements for classifying areas according to their
degree of hazard and for selecting equipment for the various areas are covered in API RP 500B* ; Section 500
of the Natl. Electrical Code19; and USCG Regulations
46 CFR, Chapter I, Subchapter I-A, Part 108 and SubParts 110-l 13. *
chapter J, Electrical Engineering,
Certifying agencies outside the U.S. have some different
rules and categories, but in general, they are no more strict
than the U.S. rules for the same situation.
Wiring Methods

and Equipment

Enclosures

Several wiring methods are applicable to offshore installations, and opinions differ as to whether conduit-and-wire
or cable-and-cable-tray
systems are best. Each has advantages and disadvantages. If the installation complies with
the appropriate procedures as outlined in the Natl. Electrical Code I9 and API RP 14F,20 each method is safe
and reliable. Rigid steel conduit and wire systems provide the maximum mechanical integrity, but the conduit
and fittings should be coated with plastic (PVC) to eliminate corrosion effectively. PVC-coated fittings and accessories are readily available in most locations. Copper-free
aluminum conduit systems have been used successfully
in lieu of galvanized steel. Aluminum systems require special wrapping of the conduit wherever steel support clamps
are used to prevent setting up corrosion cells where dissimilar metals would come in contact. Overall, conduit
systems provide excellent protection for wiring systems,
but they have the disadvantage of being bulky and relatively expensive to install, maintain, and modify.

ENGINEERING

HANDBOOK

Type MC (metal-clad) cable, which has a corrugated


aluminum sheath and an overall PVC jacket, is another
preferred system that is frequently used offshore. Armored shipboard cable and type TC tray cable are viable
alternatives to MC cable in many instances and are suitable as long as the installation complies with the Natl. Electric Code. I9 Cable systems are more flexible, quicker to
install, and less subject to corrosion than conduit offshore,
but they are more subject to mechanical damage.
As with onshore facilities, all electrical equipment installed in locations classified as Class 1, Div. 1 must be
explosion-proof.
Requirements for equipment in Class I,
Div. 2 areas are slightly less stringent as long as no arcing contacts are exposed. Electrical equipment in nonhazardous areas generally is chosen for its applicability
to the situation. Switches, for instance, generally are
provided with explosion-proof
enclosures for mechanical protection and durability even if the contacts are enclosed. General-purpose
enclosures are normally used in
protected areas. Fiberglass enclosures are seldom used
in open areas because of their susceptibility to mechanical impact damage.
Explosion-proof
motors are required in Div. 1 areas.
Fractional horsepower motors in Divs. 1 or 2 areas must
be explosion-proof.
Otherwise, both Div. 2 and unclassified areas permit the use of totally enclosed fan-cooled
(TEFC), totally enclosed nonventilated (TENV), or encapsulated, open drip-proof motors. Choice of enclosure
must be based on exposure and service. Motors operating at 2,400 V or higher should be equipped with integral
heaters or low-voltage winding heating systems if they
are in exposed locations to ensure that the integrity of the
insulation resistance is maintained
during periods of
nonuse. As warm windings cool in the relatively highhumidity offshore atmosphere, moisture is pulled into the
windings, giving rise to a high chance of an internal short
circuit. Motors of 500 hp or more that have been shut
down for an extended period should always be checked
with an insulation tester before they are restarted, regardless of whether they have been heated in the interim.

General Instrumentation
The primary guide for the instrumentation of offshore installations is API RP 14C. 24 In conjunction with the applicable U.S. OCS Orders, 2s,26 the API guide provides
an excellent reference for guidelines in the design of monitoring and control systems for offshore-production
facilities. Most offshore installations have local control
panels for packaged equipment, such as gas compressors,
low-temperature gas separation (LTS) units, and electrical generators. Beyond this, the choice of individual local display/controllers
or centralized control rooms with
field transmitters and local displays depends on the nature of the facility, the complexity of the operation, economics, and preference of the operator. Most large,
modern offshore facilities are planned with centralized
controls. The one design principle that must be followed
regardless of the overall design philosophy is that all systems should be designed to be fail-safe so that loss of a
signal represents an alarm or shutdown condition.
Even when a centralized control room has been provided, continuous remote, closed-loop (analog) control is seldom used. Remote, closed-loop controls usually are

OFFSHORE

OPERATIONS

considered only when the process includes complex separation processes, sulfur removal and handling, or gas
processing. Most oil/gas handling processes are sufticiently simple and straightforward that local control loops are
satisfactory.
Centralized controls can be based on conventional instrument and relay control panels, but advances in electronics
have increased
the use of programmable
controllers and microprocessor-based
instrument systems.
These advanced system designs have multiplexed monitoring and control systems that frequently offer significant savings in wiring costs and space for a complex
installation. They are more compact and much more flexible than conventional control panels.
Offshore instrumentation
is very similar to onshore instrumentation. Process controls frequently include a combination of pneumatic, hydraulic, electric, and electronic
instrumentation. Process variables that must be monitored
and/or controlled include level, pressure, temperature,
flow, oil/water interface, and gas/oil interface. In most
offshore installations, the considerations in designing instrument systems to accommodate these variables are identical to those for their land-based equivalent. Foam, gas
cutting, sand, wax, and HzS are typical considerations.
Radios are used extensively for offshore communications,
and radio interference can present some unique problems.
Some electronic-based
sensors are sensitive to radiofrequency interference and may give false alarms or shutdown signals when high-powered radios are keyed nearby.
Most operators prefer to separate emergency shutdown
circuits from alarm and control circuits. Typical examples are high- and low-level shutdowns on tanks or separators, high/low pilots on manifolds, and gas and fire
detectors.
Regardless of the level of complexity of an instrument
and control system, a conscientious, well-disciplined, and
well-documented
program of regular testing and maintenance is essential. Offshore instrument systems are not
overly complicated, and given adequate maintenance and
correct initial installation, every instrument should perform its intended function reliably over the life of the facility.
Safety Systems
Separation of process-related shutdowns from instrumentcontrol loops was mentioned earlier. Process shutdowns
generally involve process-related variables that have exceeded preset limits. Other key safety-related systems
should be kept separate. These systems are covered in various U.S. Coast Guard Regulations and OCS Orders. 26
Examples of these systems are combustible-gas detectors,
poisonous-gas sensors, fire detectors, surface-controlled
subsurface safety valves (SCSSVs), surface safety valves
(SSVs), and emergency shutdowns (ESDs). Proper design and installation of these systems is perhaps the single most critical aspect of offshore instrumentation
and
control systems. (Safety shut-in systems are covered in
more detail in Chap. 3).
Combustible-gas detection systems are required in most
offshore operations. They are intended as early warming
devices to alert operators to potentially hazardous conditions where none normally exist. Modern systems have
catalytic sensing heads whose electrical characteristics depend on the concentration of hydrocarbon gases surround-

18-47

ing the sensor. Units are calibrated and have known


variations in their electrical output that are based on the
gas concentration in the area of the sensor. Gas sensors
normally are connected to a central monitoring panel
equipped with individual sensor readouts calibrated in percent of lower explosive limit (LEL). Each readout has at
least two alarm outputs. Normally, one alarm is set for
about 20% LEL and the other for 60% LEL. Some operations/regulations
require automatic shut-in at the higher
level. Although gas detection technology has improved
over the years, malfunctioning caused by poisoning of the
sensing heads by contaminants in the air and loss of circulation caused by dirt accumulations on the sensors continues to be a problem.
H 2 S gas detectors are essential where sour crude and
sour gas is handled or produced because of the extreme
toxicity of Hz S. Sensors continue to improve, but reliability and maintenance are continuing problems.
Because of their vulnerability,
gas detectors must be
installed where they are protected from water spray, drilling mud, and other contaminants; yet they must be in areas
where they can adequately monitor the environment. They
generally are installed in areas where leaks or accumulations might be expected under abnormal conditionsabove gas compressors, in a wellhead or manifold area,
over drilling mud pits, or in dead-air spaces-or
in areas
where a gas build-up could be catastrophic, such as in
ventilation system inlets.
Fire-detection systems generally are based on the use
of ultraviolet (UV) sensors or fusible plugs. The operating principle of UV sensors is that their sensitivity to the
UV radiations from flame provides an alarm output in the
presence of UV radiations from open flames. Unfortunately, they also are somewhat sensitive to direct and reflected UV radiation from welding arcs. Because of their
extreme sensitivity, most UV fire systems include a brief
time delay to minimize false triggering of a fire alarm.
The layout of UV sensors at a site is important. Considerable care must be taken in laying out a coverage that considers viewing angles, range, and sensitivity.
Infrared (IR) fire sensors were tried on early offshore
platforms, but they had many operating problems. IR sensors are seldom used today, primarily because of diffculties with their calibration and reliability.
Fusible plug fields that consist of pressurized stainless
steel or plastic tubing, heat-sensitive
solder plugs, and
pneumatically held pilot shutdown valves are popular and
reliable systems for fire detection. With this system, tubing is run through and around various critical areas of
an offshore facility. The tubing runs are segregated by
process area or some other criterion. Multiple solder plugs
are included within each field. If a fire occurs in that field,
the solder plug melts, depressurizing
the field and tripping a shutdown valve. The major problem with plug
fields is maintenance and accidental shutdown because of
leaks.
Overall, a judicious combination of strategically placed
UV sensors and fusible plugs forms the optimum fire detection and automatic shut-in system. With such a SYStern, it is possible to shut in production facilities, to blow
down pressurized vessels, and to activate the appropriate
fire suppression system simultaneously
and immediately, thus minimizing fire danger.

PETROLEUM

18-48

SCSSVs should form an integral part of every offshore


production system. These valves are installed in the production tubing below the mudline. They are hydraulically actuated and held open during normal operation by
pressurized hydraulic fluid in their individual control lines.
They are designed to be fail-safe in that they close when
a loss of hydraulic pressure occurs. On modem platforms,
the SCSSV hydraulic system is generally a separate, centralized, hydraulic power unit dedicated solely to their
control. Surface power units and their associated alarms
and controls are available as specialty packaged units.
Their design and component selections have been developed over the years to the point where it is not costeffective to try to design alternative units.
Platform wells also are equipped with surface safety
valves (SSVs) between the wellhead and the production
manifold. The actuators on these valves are designed to
be fail-safe closed and can be actuated either hydraulically or pneumatically by the automatic safety shutdown
system and/or the manual emergency shutdown.
SCSSVs normally are actuated only in extreme emergency to preserve their integrity. Repairs are expensive.
SSVs, on the other hand, usually are activated in almost
all platform or process shut-ins. They are simpler devices
that are less expensive to repair, more rugged, and more
accessible.
Nearly all platforms are equipped with automatically
controlled riser shutoff valves on pipelines and flowlines
feeding or leaving the platform. The intent of these valves
is to allow isolation of the platform from any outside
source of flammable fuel in the event of a platform accident. In the case of subsea wells, it frequently is desirable to shut in the wells by simply blocking the flowline
rather than operating the subsea valves unnecessarily. The
flowline riser valves commonly are operated first in the
event of a subsea well shut-in and last on startup to avoid
cutting out seafloor valves by closing or opening them
unnecessarily against a flowing stream. Riser valves are
far more accessible and maintainable than the subsea tree
valves.
The final element in any safety system should be the
manual ESD. These controls can be either electrically
operated solenoids or pneumatically or hydraulically held
pilot valves that control various shutdown control circuits
around the facility. ESD stations usually are located on
boat landings, on helidecks, in process areas, and in control rooms.

Control of Subsea Production

Facilities

The mechanics of subsea production systems, such as


wells and manifold centers, were covered under Production Facilities. This section presents various operating
philosophies and discusses methods of providing remote
control of subsea equipment.
Subsea controls should be as simple and straightforward
as possible and still meet requirements for operational considerations and the physical layout of the field. System
reliability, maintainability,
control response times, and
the need for feedback of tubing or annulus pressure and
valve position to the operator are some of the most important factors that must be considered. Physical aspects
of the oilfield-water
depth, lateral offset of the subsea
equipment from its associated platform, anticipated sea
or ice conditions, and the potential for well damage caused

ENGINEERING

HANDBOOK

by shut-ins-and
the complexity of the subsea facility determine the optimal design of the control system.
Subsea facilities require two separate sets of controls.
Production controls provide day-to-day operational control of the subsea equipment. Initial completion and subsequent workover of subsea wells require controls
designed specifically for installation/maintenance
functions. Completion/workover
systems generally provide
control over more functions than the associated production system. Production controls normally exclude control over hydraulic connectors, test ports, and vertical
access valves.
Reliability/Maintainability
Reliability of a control system can be considered to be
the probability that the control will not malfunction in such
a manner as to preclude performing an intended function.
Reliability usually is quoted over some specified period,
such as the intended operating life of a particular project
or the planned time between scheduled maintenance. For
any system, the probability that the system will fail to perform its intended functions over a given time period is
a function of the design, quality of the system components,
built-in redundancy,
and the quality of manufacture.
Regardless of how well a system is designed and built and
how short an operating period is considered, there will
be a finite probability that it may malfunction during that
period.
The likelihood of a failure can be minimized, but it can
never be completely eliminated. Consequently, rather than
going to great lengths and expense to design a system that
cant fail, it may be more cost-effective to set realistic reliability goals and to concentrate on designing the
equipment for maximum maintainability.
This approach
is aimed at minimizing the effect of a problem instead of
trying to avoid the inevitable and generally results in a
sound design that maximizes ease of retrieval and reinstallation. How best to implement increased reliability and
maintainability
should be determined by the incremental
cost involved and the benefits to be derived.
Redundancy
A system fault generally results in a well shut-in, and
recovery of the control equipment for repair and reinstallation is then required before the well can be returned to
production. The ultimate cost of the repairs is reflected
in the expense of mobilization for repair, the repair, and
subsequent demobilization.
One of the most economic and effective means of increasing system reliability is to include active backup for
weaker elements within the basic subsea control module
or, in the extreme, to provide a completely redundant
module.
This approach may be cost-effective in situations where
field conditions could impose severe economic penalties
or pose a safety hazard in the event of a control system
malfunction. The value of deferred/lost production and
the possibility of premature well work necessitated by the
shut-in also must be taken into account in evaluating the
economics of providing redundancy.
Redundancy does not eliminate the ultimate need for
repairs, but it may permit postponing them to take advantage of favorable weather windows, contracting rate
trends, or vessel availabilities,
all of which can work to

OFFSHORE

OPERATIONS

substantially
reduce repair costs. In addition, built-in
redundancy allows operations personnel to schedule the
work on the basis of convenience without incurring unnecessary production losses or potential well damage.
Although potentially economical,
redundancy is not
free. The decision on the extent to which it should be included in a design must be based on a careful examination of the cost impact on the project and the potential
benefits.
Operational

Considerations

Many operational factors and operating philosophies must


be weighed when the control system for a specific installation is selected. Among the items that should be considered are response-time requirements; potential need for
diver or remote-operated
vehicle intervention;
requirements for feedback regarding subsea valve operation,
wellhead pressures, and temperatures; single-well or multiwell completions; use of subsea manifolds, commingling, and application of subsea chokes; type of control
fluid; and the type of control-system
design.
Response time for opening a tree valve usually is not
critical if it is not unreasonably long, particularly with
low flow velocities and clean production. Closing times
are of considerable
concern,
however, and closing
response may determine the type of control system to be
employed. Excessively long closing times are neither desirable nor necessary. From the standpoint of pollution
potential in the event of a flowline rupture, closing
response times are less critical for water-injection
wells
than clean-gas wells, which, in turn, are less critical than
oil wells.
All other factors being equal, a system that provides
discrete control over each subsea valve and allows the
operator to verify or to infer valve operations would always be recommended over one that does not. Some type
of feedback on valve position is almost mandatory for safe
operation of subsea facilities. If there is any concern about
wellhead temperature or pressure because of flowline materials or other considerations, the ability to sense this information and to transmit it to the surface can be an
overriding consideration
in the selection of a control
system.
Control Fluids
High-pressure control fluid is the means of converting a
control command into subsea valve operation in both allhydraulic and electrohydraulic control systems. Militarygrade, low-viscosity, conventional oil-based hydraulic oil
and highly water-based fluids are the two types of fluids
in subsea control systems. Oil-based fluids provide the
best system performance from the standpoints of lubricity, component wear, internal leakage, corrosion protection, and ultimately, system reliability. Oil can be used
only in closed systems, however, because it cannot be discharged into the ocean when a control loop is vented to
deactivate a control. Closed systems imply higher costs
and, over long distances, slower response times. Oil-based
systems can be particularly troublesome in cold climates.
Water-based fluids, on the other hand, are inexpensive,
are biodegradable (so they can be discharged into the environment), exhibit very low viscosity, and provide the
fastest response times.
Unfortunately, water-based fluids also have certain inherent deficiencies.
They exhibit lower lubricity, en-

18-49

courage higher leak rates, display lower corrosion


inhibition, and are subject to biofouling from bacterial
growth. All of these shortcomings can be overcome with
proper component selection, design, and operating practices. Water-based fluids are used in most subsea and drilling control systems.
The single most important factor with any hydraulic
control system, regardless of the type of fluid, is fluid
cleanliness. Failure to keep the hydraulic system clean
virtually guarantees an early malfunction.
Umbilicals
Platform control of subsea facilities requires a control umbilical between the platform and the subsea facility. The
umbilical may consist of multiple hydraulic lines in a common jacket, a composite of hydraulic hoses, power wires,
and communication pairs in a common jacket, or separate
hydraulic and electrical bundles. The makeup of the umbilical depends entirely on the nature of the control system and the field conditions.
Hydraulic bundles can be fabricated with either stainless steel tubes or elastomeric hoses. Because of the importance of quick response and because hose expansion
is a major determining factor in response time, manufacturers have upgraded hoses to the point that their expansion characteristics approach that of steel tubing. Hose
and tubing display different installation and operational
characteristics,
however, and the decision on umbilical
design must consider the unique aspects of the materials
and specific project requirements.
Protection of control lines laid on the ocean floor is always a concern. One of the principal means of protecting control lines that are not buried is to armor them with
galvanized wires applied in two separate, contrahelically
wound layers. Because of the cost of armoring the large
diameters encountered in most hose bundles, however,
they are seldom armored. As armoring costs are reduced
with improved technology, armoring of hose bundles may
become more common. Hose bundles that are armored
usually have a polyethylene bedding jacket under the armor and may have a thin covering over the armor. Some
unarmored designs include a small-diameter
wire rope
molded integrally with the jacket to provide tensile
strength and additional weighting. An alternative hosebundle design makes use of a thick outer urethane jacket
in lieu of armor. The high density of urethane provides
the necessary negative buoyancy, and it provides excellent abrasion resistance.
A viable alternative to hose bundles uses stainless steel
tubes laid side-by-side in a flat configuration.
Relatively
large-diameter wirelines (ropes) normally are placed on
the outside of a flat bundle to provide mechanical protection and resistance to kinking.
The generally accepted practice is to armor subsea electric control cable for mechanical protection, weighting,
and tensile strength. Even though it means a substantial
reduction in cost, few, if any, unarmored cables have been
installed subsea.
Alternative

Subsea Control System

There are currently two primary approaches to the control of subsea equipment-all
hydraulic and a hybrid of
electric and hydraulic (electrohydraulic).
Each offers a
number of variations with unique advantages, disadvantages, and associated costs that must be considered and

18-50

PETROLEUM

ENGINEERING

HANDBOOK

SUPPLY

CONTROLLER

SUBSEATREE
AND

Fig. 18.49-Direct

hydraulic

evaluated before a system design is finalized. Usually,


one design will prove superior to the others for a given
situation. Applicability of a given design depends on factors such as water depth, operating environment,
offset
distances, equipment to be controlled, operating requirements, operating philosophies, reservoir characteristics,
and field economics.
The following sections include brief discussions of
several basic system designs and their primary advantages
and disadvantages.
When the final selection is made,
trade-offs must be made between simplicity, response
time, operability, and costs on a site/project-specific basis.
Direct Hydraulic

Control

Direct hydraulic control (Fig. 18.49) is the most straightforward design approach. It uses a single three-way surface control valve-a
single, relatively large-diameter,
dedicated high-pressure control line between the surface
control valve and the subsea tree hydraulic-valve
actuator and the valve actuator. When the surface control valve
is operated, high-pressure fluid is introduced into the control hose, causing the subsea valve actuator to open the
tree valve. When the surface valve is deactivated, the fluid
that opened the subsea valve is returned to the surface
fluid reservoir. The advantages of this system are simplicity, discrete remote control over each subsea function,

subsea control.

inherent feedback on subsea operations, and minimum


cost of the basic control hardware. The disadvantages are
slowest response time for a given control-line size, highest
control-line costs, and potential problems with corrosion
and biofouling of the subsea actuator when a water-based
control fluid is used because the fluid is never renewed.
System costs are directly proportional to the number of
functions controlled and the distance between the control
point and the subsea device. This type of system is seldom used with producing wells at offset distances of more
than 10,000 ft, or with injection wells at more than 15,ooO
ft, because of the cost and response-time considerations.
Response time of the direct hydraulic controls for the
valve-closing (shut-in) operation over a given distance
with a given line size can be improved significantly by
installing a subsea dump valve on each tree-valve actuator. This approach also permits the introduction of new
control fluid into the control lines because the fluid in the
actuators is vented to the ocean whenever a valve is closed.
This provides a gradual renewal of control fluid with each
valve operation.
The disadvantage of dump valves is that they introduce
additional flow-rate/direction-sensitive
hydraulic-control
elements that require very clean control fluid. This increases subsea hardware cost and markedly reduces the
overall system reliability.

/
COMMON

VALVE

ACTUATOR

H P SUPPLY

PILOT

VALVE

(1 PER FUNCTION)

1
I

SUBSEA
AND

k------i
I
I

Fig. 18.50-Discrete-piloted

hydraulic

subsea control

TREE

ACTUATOR

VALVE

OFFSHORE

OPERATIONS

18-51

Fig. l&51--Sequential-piloted

Discrete-Piloted

Hydraulic

Discrete-piloted control (Fig. 18.50) has a single threeway control valve at the surface for each subsea function,
a corresponding
subsea pilot control valve, and single
small-diameter dedicated signal line between the two control valves. A single common high-pressure line provides
hydraulic supply from the surface to the tree. When the
subsea pilot valve is actuated, it switches high-pressure
fluid from the supply line to the subsea valve actuator.
When the pilot valve is deactuated, it vents the actuator
fluid subsea rather than returning it to the surface.
The advantages of this design are its relative simplicity, discrete remote control over each subsea function, inherent inferred feedback on subsea operations, relatively
low-cost hardware, and relatively fast response compared
to the direct hydraulic design. The disadvantages are that
cost and response time are both proportional to distance
just as they are with direct hydraulics, but the effects are
significantly less. The pilot lines are dead-ended, but cor-

hydraulic

subsea control.

rosion is not a problem because pilot valves usually are


made of stainless alloys. Biofouling, however, can still
be a problem.
Sequential-Piloted

Hydraulic

Sequential control is similar to discrete-piloted control in


that it also has subsea pilot valves that direct high-pressure
hydraulic fluid to tree-valve actuators (Fig. 18.5 1). However, subsea pressure-sensitive
pilot valves that are
manifolded to a common signal line are used in this design. Rather than discrete control over each individual pilot valve, they are switched in groups according to the
signal pressure and the pilot-valve set points. The valves
are interconnected
so that supply pressure is applied to
the subsea tree actuators in a predetermined sequence in
response to changes in signal pressure. Up to six combinations of valve operations can be implemented reliably
with a single control line and a single supply line. The
advantage of this system is its relatively lower cost for

1 ATMOS

SlNGLE WIRE PAIR REGARDLESS


OF NUMBER OF SIGNALS
OR
AMOUNT OR TYPE OF INFORMATION

POD

1 ATMOS
AMBIENT

SOLENOID
VALVE \

-I----

COMMON

ELECTRICAL

POWER

OR OIL FILLED
COMP
POD

SUPPLY

ACCUMULATOR

COMMON

H P HYDRAULIC

SUPPLY

SUBSEATREE
VALVE WITH
HYDRAULIC
ACTUATOR

Fig. l&52-Multiplexed

control with hydraulic

valve actuators.

1 B-52

PETROLEUM

hardware and control lines. It has a somewhat slower


response than a discrete-piloted hydraulic system, and it
is less reliable because it is more complex. The really significant disadvantages,
however, are that independent
valve control is impossible, there is no ready means of
confirming valve operations, and operating sequences
must be predetermined before manufacture and installation. Further, the sequence approach offers no means for
independent
valve control for performing operational
tests, troubleshooting or diagnosis, or for changing flow
paths through a subsea system in the event of a malfunction.
Multiplexed

Electrohydraulic

Control

This design (Fig. 18.52) has a single high-pressure


hydraulic supply line and an electric cable that, in its simplest form, consists of one pair of signalling wires and
one pair of power wires. The limited number of electrical conductors permits the use of inductive couplers rather
than the more conventional pin-and-socket connectors. Pin
connectors historically have been troublesome in subsea
production applications,
while inductive couplers have
proved to be relatively foolproof. Inductive couplers are
easily and reliably remotely mated and unmated subsea,
which provides maximum flexibility in hardware layout.
Encoding and decoding multiplex logic are provided at
the surface and subsea to enable transmission of coded
valve-open and valve-close commands and associated status feedback over one pair of single wires. The multiplexed
valve commands trigger electric solenoid pilot valves that
activate hydraulic-switching
valves to direct high-pressure
fluid to the tree-valve actuators. The advantages of this
system are (1) fastest possible response independent of
distance, (2) simplicity of control lines, and (3) total flexibility. With some designs, a totally hydraulic backup system can be included with the multiplexed system at little
increase in cost to provide some measure of temporary
operability in the event of an electrical failure. The obvious disadvantages of this design are its increased complexity and higher costs.
With a multiplexed electrohydraulic
system, a recirculating (closed) subsea hydraulic power unit is also feasible in the control housing. This approach offers the
possibility of eliminating the hydraulic supply line if system leakage can be controlled to a very low level. If this
design approach is used, some provision still must be
made for fluid makeup to account for long-term leakage.
The primary advantage is that it allows the use of conventional oil-based hydraulic fluids, thus ensuring maximum life and reliability. The obvious disadvantages are
the substantial increase in system complexity and consequent overall lower system reliability.

References
I. Rintoul, B.: Drilling
From The Steel Island,
Pcrcific Oi/
World/Annual, Petroleum Publishers, Brea. CA (Jan. 1980) 73,
18-28.

ENGINEERING

HANDBOOK

2. The Tt4~nolog~ ojQfj~hore Drilling, Coi,~pl~tiotc md Pmductim.


ETA Offshore Seminars, Inc.. The Petroleum Publishing Co.. Tulsa
(1976).
3. Silcox, W.H.: Floating Drilling: The First 30 Years--Part
I and
Part 2, J. Per. Tech. (Jan. and Feb. 1983).
4. Burke, B.G.: Downtime Evaluation for Operations from Floating Vessels in Waves, Proc., 1977 Society of Naval Architects
and Marine Engineers Second Ship Tech. and Research Symposium. New York City, 279-301.
5. Sheffield, R.: Floating Drilling: Equipment and Its Use, Prrw
tica/ Drilling Technology. Gulf Publishing Co., Houston (1980) 2.
6. Recommended
Practice for Blowout Prevention Equipment Systerns, second edition, API RP 53. API, Dallas (Jan. 1984).
7. The Analysis of Spread Mooring Systems for Floating Drilling
Units, latest edition, API RP 2P. API, Dallas.
8. Design and Operation of Marine Drilling Riser Systems. latest
edition, API RP 24, API, Dallas.
9. Recommended
Practice for Care and Use of Marine Drilling
Risers, latest edition. API RP 2K. API. Dallas.
IO. Recommended
Practices for Safe Drilling of Wells Contaimng
Hydrogen Sulfide. latest edition, API RP 49, API, Dallas.
I I. Strengthening for Navigation in Ice. Rulrsfor Buikditlg trnd Clmsing Steel Vessels, latest edition. American Bureau of Shipping, New
York City.
12. Bardgette, J.J. and Irick, J.T.: Construction
of the Hondo Platform in 850 Feet of Water, paper OTC 2959 presented at the 1977
Offshore Technology Conference, Houston. May 2-S.
13. Kinra, R.K. and Marshall, P.W.: Fatigue Analysis of the Cognac
Platform, paper OTC 3378 presented at the 1979 Offshore Technology Confeience, Houston. April 30-May 3.
14. Tan&ill, C.A., Isenhower, W.M., and Engle. D.D.: Cerveza-A
Project Overview of a Deep-water
Platf&m,
paper OTC 4185
presented at the 1982 Offshore Technology Conference, Houston.
May 3-6.
15. Goodman, M.A.: Handbook of Arciic Well Completions. Gulf Publishing Co., Houston (1978).
16. U.S. Arctic Oil and Gas. Natl. Petroleum Council, Washington, DC (Dec. 1981).
17. API Bulletin on Planning, Designing, and Constructing Fixed Offshore Structures in Ice Environments,
latest edition, Bull. 2N.
API, Dallas.
18. Willits, K.L. and Lindsey, W.C.: Well Completions in the Prudhoe
Bay Field, Pet. Eng. Id. (Feb. 1976) 48-56.
19. Narl. Elecrrical Code, latest edition, Natl. Fire Protection Assn..
Quincy, MA.
20. Recommended
Practice for Design and Installation of Electrical
Systems for Offshore Production Platforms,
latest edition, API
RP 14F, API, Dallas.
21. USCG Regulation 46 CFR Shipping, Chap. I, Subchapter I-A, MObile Offshore Drilling Units, Subchapter J, Electrical Engineering, U.S. Government Printing Office, Washington. DC. Parts
108 and 110-13.
22. Recommended
Practice for Classification of Areas for Electrical
Installations at Drilling Rigs and Production Facilities on Land and
on Marine Fixed and Mobile Platforms,
latest edition, API RP
5OOB, API. Dallas.
23. Standards for Purged and Pressurized Enclosures for Electrical
Equipment in Hazardous (Classified) Locations, Bull. 496, Natl.
Fire Protection Assn., Quincy, MA.
24. Recommended Practice for Analysis, Design, Installation and Tcsting of Basic Surface Safety Systems on Offshore Production Platforms, latest edition, API RP 14C, API, Dallas.
25. LJSCG Regulation 33 CFR, Chap. I, Subchapter N. Outer Continental Shelf Activities,
U.S. Ciovernmknt Printing Office,
Washington. DC, Parts 140 through 147.
26. USCG Regulation 30 CFR, Part 250, Oil and Gas and Sulfur Operations in the Outer Continental Shelf. Minerals Management
Service, U.S. Dept. of Interior, Washington,
D.C.

Das könnte Ihnen auch gefallen