0% found this document useful (0 votes)
46 views120 pages

Lecture Notes - Production Engineering

Uploaded by

nardelee550
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
46 views120 pages

Lecture Notes - Production Engineering

Uploaded by

nardelee550
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 120

Advanced Production Engineering 2019/2020 Academic Year

CHAPTER 1
HISTORY OF OIL AND GAS PRODUCTION

1.1 INTRODUCTION

Oil has been used for lighting purposes for many thousand years. In areas where oil is
found in shallow reservoirs, seeps of crude oil or gas may naturally develop, and some oil
could simply be collected from seepage or tar ponds. Historically, we know of tales of
eternal fires where oil and gas seeps would ignite and burn. One example 1000 B.C. is the
site where the famous oracle of Delphi would be built, and 500 B.C. Chinese were using
natural gas to boil water.

But it was not until 1859 that "Colonel" Edwin Drake drilled the first successful oil well,
for the sole purpose of finding oil. The Drake Well was located in the middle of quiet farm
country in north-western Pennsylvania, and began the international search for and
industrial use of petroleum.

Figure 1.1: Drake well museum collection, Titusville, PA

These wells were shallow by modern standards, often less than 50 meters, but could give
quite large production. In October 1861, Phillips well was drilled on the right of the Oil
Creek Valley and was initially flowing at 4000 barrels per day whereas and the Woodford
well on the left of the Oil Creek Valley came in at 1500 barrels per day in July, 1862. The
oil was collected in the wooden tank in the foreground. Note the many different sized
barrels in the background. At this time, barrel size was not yet standardized, which made
terms like "Oil is selling at $5 per barrel" very confusing (today a barrel is 159 liters). But
even in those days, overproduction was an issue to be avoided. When the “Empire well”
was completed in September 1861, it gave 3,000 barrels per day, flooding the market, and
the price of oil plummeted to 10 cents a barrel.

Soon, oil had replaced most other fuels for mobile use. The automobile industry developed
at the end of the 19th century, and quickly adopted the fuel. Gasoline engines were

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 1


Advanced Production Engineering 2019/2020 Academic Year

essential for designing successful aircraft. Ships driven by oil could move up to twice as
fast as their coal fired counterparts, a vital military advantage. Gas was burned off or left
in the ground.

1.2 OVERVIEW OF OIL AND GAS PRODUCTION

The following figure gives a simplified overview of the typical oil and gas production
process.

Figure 1.2: Oil and gas production overview

At the left side, we find the wellheads. They feed into production and test manifolds. In a
distributed production system this would be called the gathering system. The remainder
of the figure is the actual process, often called the Gas Oil Separation Plant (GOSP). When
there are both oil or gas, more often the wellstream will consist of a full range of
hydrocarbons from gas (methane, butane, propane etc.), condensates (medium density
hydrocarbons) to crude oil. With this well flow we will also get a variety of non-wanted
components such as water, carbon dioxide, salts, sulfur and sand. The purpose of the
GOSP is to process the well flow into clean marketable products: oil, natural gas or
condensates. Also included are a number of utility systems, not part of the actual process,
but providing energy, water, air or some other utility to the plant.
Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 2
Advanced Production Engineering 2019/2020 Academic Year

1.2.1 MAIN PROCESS SECTIONS

WELLHEADS
A wellhead consists of the pieces of equipment mounted at the opening of the well to
regulate and monitor the extraction of hydrocarbons from the underground formation.
The wellhead sits on top of the actual oil or gas well leading down to the reservoir. A
wellhead may also be an injection well, used to inject water or gas back into the reservoir
to maintain pressure and levels to maximize production. Once a natural gas or oil well is
drilled, and it has been verified that commercially viable quantities of natural gas are
present for extraction, the well must be 'completed' to allow for the flow of petroleum or
natural gas out of the formation and up to the surface. This process includes strengthening
the well hole with casing, evaluating the pressure and temperature of the formation, and
then installing the proper equipment to ensure an efficient flow of natural gas out of the
well. The well flow is controlled with a choke. The wellhead structure, often called a
Christmas tree, must allow for a number of operations relating to production and well
workover. Well workover refers to various technologies for maintaining the well and
improving its production capacity.

Figure 1.3: Wellhead

Wellheads can be Dry or Subsea (Wet) completion. Dry Completion means that the well is
onshore on the topside structure on an offshore installation. Subsea (Wet) wellheads are
located under water on a special sea bed template that allows the wells to be drilled and
serviced remotely from the surface. Wellhead prevents leaking of oil or natural gas out of
the well, and prevents blowouts due to high pressure formations. Formations that are
under high pressure typically require wellheads that can withstand a great deal of upward
pressure from the escaping gases and liquids. These wellheads must be able to withstand
pressures of up to 140 MPa (1400 Bar). The wellhead consists of three components: the
casing head, the tubing head, and the ‘Christmas tree’. A typical Christmas tree is
composed of a master gate valve, a pressure gauge, a wing valve, a swab valve and a choke
as shown in Figure 1.4.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 3


Advanced Production Engineering 2019/2020 Academic Year

Figure 1.4: Christmas tree

The Christmas tree may also have a number of check valves. The functions of these
devices are explained in the following paragraphs.

At the bottom we find the Casing Head and Casing Hangers. The casing will be screwed,
bolted or welded to the hanger. Several valves and plugs will normally be fitted to give
access to the casing. This will permit the casing to be opened, closed, bled down, and, in
some cases, allow the flowing well to be produced through the casing as well as the tubing.
The valve can be used to determine leaks in casing, tubing or the packer, and will also be
used for lift gas injection into the casing.

The tubing hanger (also called donut) is used to position the tubing correctly in the well.
Sealing also allows Christmas tree removal with pressure in the casing.

The master gate valve is a high quality valve. It will provide full opening, which means
that it opens to the same inside diameter as the tubing so that specialized tools may be
run through it. It must be capable of holding the full pressure of the well safely for all
anticipated purposes. This valve is usually left fully open and is not used to control flow.

The pressure gauge. The minimum instrumentation is a pressure gauge placed above the
master gate valve before the wing valve to check for pressure of the flow. In addition other
instruments such as temperature will normally be fitted.

The wing valve. The wing valve can be a gate valve, or ball valve. When shutting in the
well, the wing gate or valve is normally used so that the tubing pressure can be easily
read.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 4


Advanced Production Engineering 2019/2020 Academic Year

The swab valve. The swab valve is used to gain access to the well for wireline operations,
intervention and other workover procedures, on top of it is a tree adapter and cap that
will mate with various equipment.

The variable flow choke valve. The variable flow choke valve is typically a large needle
valve. Its calibrated opening is adjustable in 1/64 inch increments (called beans). High-
quality steel is used in order to withstand the high-speed flow of abrasive materials that
pass through the choke, usually for many years, with little damage except to the dart or
seat. If a variable choke is not required, a less expensive positive choke is normally
installed on smaller wells. This has a built in restriction that limits flow when the wing
valve is fully open.

The Christmas tree in Figure 1.4 is a vertical tree. Christmas trees can also be horizontal,
where the master, wing and choke is on a horizontal axis. This reduces the height and may
allow easier intervention. Horizontal trees are especially used on subsea wells.

MANIFOLDS/GATHERING
Onshore, the individual well streams are brought into the main production facilities over a
network of gathering pipelines and manifold systems. The purpose of these is to allow set
up of production “well sets” so that for a given production level, the best reservoir
utilization, well flow composition (gas, oil, water) etc. can be selected from the available
wells. For gas gathering systems, it is common to meter the individual gathering lines into
the manifold as shown on the illustration. For multiphase (combination of gas, oil and
water) flows, the high cost of multiphase flow meters often lead to the use of software
flow rate estimators that use well test data to calculate the actual flow. Offshore, the dry
completion wells on the main field centre feed directly into production manifolds, while
outlying wellhead towers and subsea installations feed via multiphase pipelines back to
the production risers.

Figure 1.5: Manifolds

Risers are the system that allow a pipeline to “rise” up to the topside structure. For
floating or structures, this involves a way to take up weight and movement. For heavy
crude and in arctic areas, diluents and heating may be needed to reduce viscosity and
allow flow.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 5


Advanced Production Engineering 2019/2020 Academic Year

SEPARATION
Some wells have pure gas production which can be taken directly to gas treatment and/or
compression. More often, the well gives a combination of gas, oil and water and various
contaminants which must be separated and processed. The production separators come
in many forms and designs, with the classical variant being the gravity separator.
In gravity separation the well flow is fed into a horizontal vessel. The retention period is
typically 5 minutes, allowing the gas to bubble out, water to settle at the bottom and oil to
be taken out in the middle. The pressure is often reduced in several stages (high pressure
separator, low pressure separator etc.) to allow controlled separation of volatile
components. A sudden pressure reduction might allow flash vaporization leading to
instabilities and safety hazards.

Figure 1.6: Horizontal separator

GAS COMPRESSION
Gas from a pure natural gas wellhead might have sufficient pressure to feed directly into a
pipeline transport system. Gas from separators has generally lost so much pressure that it
must be recompressed to be transported. Turbine compressors gain their energy by using
up a small proportion of the natural gas that they compress. The turbine itself serves to
operate a centrifugal compressor, which contains a type of fan that compresses and
pumps the natural gas through the pipeline. Some compressor stations are operated by
using an electric motor to turn the same type of centrifugal compressor. This type of
compression does not require the use of any of the natural gas from the pipe; however it
does require a reliable source of electricity nearby. The compression and treatment
include a large section of associated equipment such as scrubbers (removing liquid
droplets) and heat exchangers, lube oil treatment etc.

Whatever the source of the natural gas, once separated from crude oil (if present) it
commonly exists in mixtures with other hydrocarbons; principally ethane, propane,
butane, and pentanes. In addition, raw natural gas contains water vapour, hydrogen
sulfide (H2S), carbon dioxide, helium, nitrogen, and other compounds. Natural gas
processing consists of separating all of the various hydrocarbons and fluids from the pure
natural gas, to produce what is known as 'pipeline quality' dry natural gas. Major
transportation pipelines usually impose restrictions on the makeup of the natural gas that
is allowed into the pipeline. That means that before the natural gas can be transported it
must be purified. Associated hydrocarbons, known as 'natural gas liquids' (NGL) are used
as raw materials for oil refineries or petrochemical plants, and as sources of energy.

METERING, STORAGE AND EXPORT


Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 6
Advanced Production Engineering 2019/2020 Academic Year

Most plants do not allow local gas storage, but oil is often stored before loading on a
vessel, such as a shuttle tanker taking the oil to a larger tanker terminal, or direct to crude
carrier. Offshore production facilities without a direct pipeline connection generally rely
on crude storage in the base or hull, to allow a shuttle tanker to offload about once a week.
A larger production complex generally has an associated tank farm terminal allowing the
storage of different grades of crude to take up changes in demand, delays in transport etc.

Figure 1.7: Metering system and oil storage

Metering stations allow operators to monitor and manage the natural gas and oil exported
from the production installation. These metering stations employ specialized meters to
measure the natural gas or oil as it flows through the pipeline, without impeding its
movement. Typically the metering installation consists of a number of meter runs so that
one meter will not have to handle the full capacity range, and associated loops so that the
meter accuracy can be tested and calibrated at regular intervals.
Pipelines can measure anywhere from 6 to 48 inches in diameter. In order to ensure the
efficient and safe operation of the pipelines, operators routinely inspect their pipelines for
corrosion and defects. This is done through the use of sophisticated pieces of equipment
known as pigs. Pigs are intelligent robotic devices that are propelled down pipelines to
evaluate the interior of the pipe. Pigs can test pipe thickness, and roundness, check for
signs of corrosion, detect minute leaks, and any other defect along the interior of the
pipeline that may either impede the flow of gas, or pose a potential safety risk for the
operation of the pipeline. Sending a pig down a pipeline is known as 'pigging' the pipeline.
The export facility must contain equipment to safely insert and retrieve pigs from the
pipeline as well as depressurization, referred to as pig launchers and pig receivers.

UTILITY SYSTEMS
Utility systems are systems which does not handle the hydrocarbon process flow, but
provides some utility to the main process safety or residents. Depending on the location of
the installation, many such functions may be available from nearby infrastructure (e.g.
electricity). But many remote installations must be fully self-sustainable and thus must
generate their own power, water etc.

1.3 OIL AND GAS PRODUCTION FACILITIES

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 7


Advanced Production Engineering 2019/2020 Academic Year

1.3.1 Onshore Facilities


Onshore production is economically viable from a few tens of barrels a day upwards. Oil
and gas are produced from several million wells world-wide. In particular, a gas gathering
network can become very large, with production from hundreds of wells, several hundred
kilometers/miles apart, feeding through a gathering network into a processing plant.

Figure 1.8: Onshore oil production

The picture shows a well equipped with a sucker rod pump (donkey pump) often
associated with onshore oil production. However, as we shall see later, there are many
other ways of extracting oil from a non-free flowing well.

For the smallest reservoirs, oil is simply collected in a holding tank and collected at
regular intervals by tanker truck or railcar to be processed at a refinery.
But onshore wells in oil rich areas, there may be also high capacity wells with thousands
of barrels per day, connected to a 1MM barrel a day gas oil separation plant (GOSP).
Product is sent from the plant by pipeline or tankers. The production may come from
many different license owners. Metering and logging of individual wellstreams into the
gathering network are important tasks.

Recently, very heavy crude, tar sands and oil shales have become economically extractible
with higher prices and new technology. Heavy crude may need heating and diluent to be
extracted. Venezuela and Canada have huge reserves of heavy and extra heavy oil. In
Canada this is called in situ crude bitumen. The major challenge of recovering bitumen
from depth is to overcome its high viscosity to allow it to flow to the wellbore. To do this,
thermal (or other nonprimary) in-situ methods are used, most commonly Cyclic Steam
Stimulation (CSS) and Steam Assisted Gravity Drainage (SAGD).

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 8


Advanced Production Engineering 2019/2020 Academic Year

Figure 1.9: Heavy oil steam process

Tar sands (upgraded and cleaner heavy oils from bitumen – synthetic crude oil) have lost
their volatile compounds and are strip mined or could be extracted with steam. The oil
sands consist of bitumen (soluble organic matter, solid at room temperature) and host
sediment, with associated minerals, and excluding any related natural gas. The crude
bitumen within the sands is a naturally occurring viscous mixture of hydrocarbons
(generally heavier than pentane), often with sulphur compounds, that will not flow to a
wellbore in its natural state. Upon heating, the bitumen will flow, and on a hot summers’
day, bitumen oozes from the outcrops along the river valleys in northeast Alberta.
Synthetic crude is a type of crude oil developed by upgrading bitumen (a tar like
substance found in tar sands). Once upgraded to synthetic crude, the oil can be
transported via pipeline for refining to gasoline and other types of petroleum products.
This is done using four main processes:

• Coking removes carbon and breaks large bitumen molecules into smaller parts.
• Distillation sorts mixtures of hydrocarbon molecules into their components
• Catalytic conversions help transform hydrocarbons into more valuable forms
• Hydrotreating is used to help remove sulphur and nitrogen and add hydrogen to
molecules

Figure 1.10: Mining of Tar Sand (Heavy Crude)

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 9


Advanced Production Engineering 2019/2020 Academic Year

Oil and gas shales – Shale consists of consolidated clay sized particles. It is the most
common sedimentary rock. It looks like slate and has almost no permeability. Shale is
usually a cap rock of many oil reservoirs. In some conditions, shale can be the source as
well as the reservoir rock and is very rich in hydrocarbons. The kerogen in oil shale can be
converted to oil through the chemical process of pyrolysis. Estonia, Russia, Brazil, China,
Australia, Israel and Germany currently mine oil shale. Depending on the thermal history
of the shale it can contain oil and gas or mainly gas (more thermally mature shales). There
are over 35,000 producing gas shale wells in the United States. The Barnett shale in Texas
is the most active gas shale exploration project in the United States. In shale reservoirs,
natural gas is stored three ways:
• As free gas within the rock pores.
• As adsorbed gas on organic material.
• As free gas within the system of natural fractures
These unconventional of reserves may contain more than double the hydrocarbons found
in conventional reservoirs

1.3.2 Offshore Facilities


In offshore production, depending on size and water depth, a whole range of different
structures are used. Offhsore production systems are usually much more complex than
onshore ones since drilling, production, treatment and storage must be done from one or
more centralized structures installed at a certain water depth as opposed to the
equivalent structures located on the ground in onshore oil fields.

Figure 1.11: Some Offshore Systems

The technology is strongly dependent on the water depth. Offshore systems are then
currently classified according to the water depth as:
• Conventional < 3000 ft
• Deep water < 5000 ft
• Ultra deep water > 5000 ft

The production systems available for offshore production can be: Fixed, Floating or Sub-sea.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 10


Advanced Production Engineering 2019/2020 Academic Year

Figure 1.12: Offshore Systems at different water depths

Almost all production systems with the exception of FPS’s (Floating Production Systems) and
FPSO’ s (Floating Platform, Storage and Offloading) can have dry trees on the platform deck.
FPS’s and FPSO’s can only use wet trees. Sub-sea systems can have single or multiple
wellheads on the sea floor connected directly or through a manifold and common production
flow line to a host platform.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 11


Advanced Production Engineering 2019/2020 Academic Year

Figure 1.13: Fixed Production Facility with Workover Rig

Figure 1.14: Floating Production Facility with Workover Rig

System type Maximum water depth (ft)


Concrete platforms 1000
Steel platforms 1500
Compliant towers 3000
Tension leg platform 5000
Spars 10000
Floating Production Systems Unlimited
Sub-sea Unlimited

Concrete platforms
Concrete platforms are large structures that sit on the seabed, stabilized by their own
massive weight. It is designed to withstand forces from the weight of the structure with
deck loads, stored oil and ballast. It must also withstand to environmental parameters as
waves, currents and winds and in some locations, earthquakes and ice loading.

Advantages:

• Oil storage capability.


• Cheaper installation cost than steel jackets since most of the equipment installation
and hook-up is performed onshore.
• Larger than steel jackets, so permit greater production and in the same time reduce
the total number of platforms required for the development of a field.
• Concrete does not require maintenance and does not suffer any reduction in load-
carrying capacity over the years.

Disadvantages:

• Relatively expensive capital cost, compared to steel jacket.


• As yet, no feasible means of removing (decommissioning) the structure has been
defined).

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 12


Advanced Production Engineering 2019/2020 Academic Year

Figure 1.15: Concrete platform

Steel platforms
A fixed steel platform consists of a welded tubular steel jacket, deck, and surface facility.
The jacket and deck make up the foundation for the surface facilities. Piles are driven into
the seafloor to secure the jacket. Once the jacket is secured and the deck is installed,
additional modules are added for drilling, production, and crew operations (topsides).
Large, barge-mounted cranes position and secure the jacket prior to the installation of the
topsides modules. After the onshore fabrication of the jacket is completed, it is loaded
onto a very large barge that will transport the jacket to its location. The towing of the
jacket may involve the use of several tugs (up to 52,000 hp combined) over hundreds of
miles. In some designs, there is a jacket base section that may be in place before the actual
jacket is installed. The placement of the jacket base section prior to the jacket could
provide better support during installation. Once the jacket arrives on location, it is
launched, up-ended, and lowered into position with two or more tugs.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 13


Advanced Production Engineering 2019/2020 Academic Year

Figure 1.16: Steel platform

With a beacon system or with a remotely operated vehicle (ROV) assist, the jacket is
placed in position on the seafloor. The beacon system consists of homing devices laid on
the seafloor around the area of the jacket’s intended site; the beacons allow computer-
aided control and monitoring of the installation process. Then a pile and hammer-
handling barge is brought in to drive the piles into the seafloor, through guides in the legs.
A second method of pile driving is the use of an underwater hammer with ROV alignment.
Once this work is completed, the jacket is secure on location and the surface facilities can
be installed.

The surface facilities are fabricated onshore and towed out on one or more crane barges.
Once on location, the crane barge(s) is moored in whatever fashion needed and
installation begins. Mooring can be done by different methods such as lines to the seafloor

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 14


Advanced Production Engineering 2019/2020 Academic Year

only or a combination of lines to the jacket and the seafloor. A crane on the barge(s)
transfers the modules as a whole or separately from the barge(s) to the deck where
workers complete the final connections.

Compliant towers
Compliant towers are much like fixed (Steel) platforms. They consist of a narrow tower
(the jacket has smaller dimensions than those of a fixed platform) and attached to a
foundation on the seafloor with piles. This tower is flexible, as opposed to the relatively
rigid legs of a fixed platform. This flexibility allows it to operate in much deeper water, as
it can 'absorb' much of the pressure exerted on it by the wind and sea. After onshore
fabrication, the jacket is towed in one or two pieces out to the site on a specially designed
barge. The process is similar to how it is done for fixed platforms. For normal compliant
towers, the upper and lower jacket can be joined at sea or vertically joined at the site. The
jacket can be launched from the rear or side of the barge. After the jacket is installed, a
deck barge brings the surface facilities out from shore to the site and installs them either
module by module or as a complete unit.

Figure 1.17: Compliant platform

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 15


Advanced Production Engineering 2019/2020 Academic Year

Tension Leg Platforms


Tension Leg Platforms (TLP) consist of a structure held in place by vertical tendons
connected to the sea floor by pile-secured templates. The structure is held in a fixed
position by tensioned tendons, to enable the use of the TLP in a broad water depth range
up to about 5000 ft. It has limited vertical motion. The tendons are constructed as hollow
high tensile strength steel pipes that carry the spare buoyancy of the structure and ensure
limited vertical motion. The foundation is the link between the seafloor and the TLP. Most
foundations are templates laid on the seafloor, then secured by concrete or steel piles
driven into the seafloor by use of a hydraulic hammer, but other designs can be used such
as a gravity foundation. The foundations are built onshore and towed to the site. The
available types are Tension Leg Platform, Tension leg wellhead platform, Mini-tension leg
platform, Mini-tension leg wellhead plat-forms.

TLP Mini-TLP

Figure 1.18: Tension Leg Platforms

SPAR
The SPAR consists of a single tall floating cylinder hull, supporting a fixed deck. The
cylinder however does not extend all the way to the seafloor, but instead is tethered to the
bottom by a series of cables and lines. The large cylinder serves to stabilize the platform in
the water, and allows for movement to absorb the force of potential hurricanes. Its four
major systems are hull, moorings, topsides, and risers. About 90 percent of the structure
is underwater. Spars can be quite large and are used for water depths from 300 and up to
3000 meters. SPAR is not an acronym, but refers to its likeness with a ship’s spar. Spars
can support dry completion wells, but is more often used with subsea wells. The hull is

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 16


Advanced Production Engineering 2019/2020 Academic Year

positioned on location by a tug and positioning system assistance. Then the mooring
system is connected to the hull. After the mooring system is connected, the lines are pre-
tensioned. Then the hull is ballasted to prepare for the topsides installation and removal
of the temporary work deck. Topsides are transported offshore on a material barge and
lifted into place by the derrick barge. An important characteristic is that the derrick barge
can perform the lift in dynamic positioning mode. There are different types of SPAR
including Classic SPAR, truss SPAR, Cell SPAR, etc.

Figure 1.19: SPAR

Floating Production Systems


An FPS system is an offshore production facility that is typically ship-shaped. They are an
amalgam of marine and petroleum functions, and therefore, present many specialized
challenges for those involved in their creation. Floating, production, storage, and
offloading systems (FPSO) receive crude oil from deepwater wells and store it in their hull
tanks until the crude can be pumped into shuttle tankers or oceangoing barges for
transport to shore. In addition to FPSO’s , there have been a number of ship-shaped
Floating Storage and Offloading (FSO) systems (vessels with no production processing

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 17


Advanced Production Engineering 2019/2020 Academic Year

equipment) used in these same areas to support oil and gas developments. An FSO is
typically used as a storage unit for processed production from other platforms that are
remote from infrastructure and lack an oil pipeline to transport the oil to the refinery. The
crude oil is periodically offloaded to shuttle tankers or oceangoing barges for transport to
shore. FPSO’s may be used as production facilities to develop marginal oil fields or fields
in deepwater areas remote from pipeline infrastructure. FPSO’s have been used to
develop offshore fields around the world since the late 1970’s. They have been used
predominately in the North Sea, Brazil, Southeast Asian/South China Seas, the
Mediterranean Sea, Australia, and off the West Coast of Africa. Their turret structures are
designed to anchor the vessel, allow “weather vaning” of the units to accommodate
environmental conditions, permit the constant flow of oil and production fluids from
vessel to undersea field, all while being a structure capable of quick disconnect in the
event of emergency.

Figure 1.20: Floating Production Systems

Subsea production systems


Subsea production systems are wells located on the sea floor, as opposed to at the surface.
Like in a floating production system, the petroleum is extracted at the seafloor, and then
can be 'tied-back' to an already existing production platform or even an onshore facility,
Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 18
Advanced Production Engineering 2019/2020 Academic Year

limited by horizontal distance or “offset”. The well is drilled by a moveable rig and the
extracted oil and natural gas is transported by undersea pipeline and riser to a processing
facility. This allows one strategically placed production platform to service many wells
over a reasonably large area. Subsea systems are typically in use at depths of 7,000 feet or
more, and do not have the ability to drill, only to extract and transport. Drilling and
completion is performed from a surface rig.

Figure 1.21: Subsea Production Systems

1.4 PETROLEUM PRODUCTION ENGINEERS


Once the well is completed, the production engineer takes over. His/her job is to analyze,
interpret, and optimize the performance of individual wells. The production engineer is
responsible for determining how to bring hydrocarbons to the surface. The production
engineer will determine the most efficient means to develop the field. The production
engineer is also responsible for the developing a system of surface equipment that will
separate the oil, gas, and water. As the field matures, the production engineer will be
responsible for exploring additional technologies to enhance production from wells that
are declining. In doing so, the production engineer will work closely with engineers in
other disciplines such as formation evaluation, drilling, and reservoir engineering, to
determine the optimal approach for that particular field.

Production engineers must have a solid background in the following technical areas:
Phase Equilibrium, Multiphase Flow in Pipes and Restrictions; Natural Flow and Artificial
Lift; Gathering and Distribution Systems; Power Generation and Distribution Systems;
Processing Surface Facilities; and Flow Assurance. To properly work in multi-disciplinary
teams, production engineers must have a good understanding of other technical areas of
Petroleum Engineering such as Drilling and Completion, Reservoir Engineering and
Enhanced Oil Recovery, Formation Evaluation, Well Testing and Well Stimulation. The
development of deep water oilfields and heavy oil production requires production
engineers to be also familiar with Sub-sea Engineering, Offshore Production Systems and
Thermal recovery methods.
CHAPTER 2

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 19


Advanced Production Engineering 2019/2020 Academic Year

PRODUCTION SYSTEM

2.1 INTRODUCTION
When oil/gas is first found in the reservoir, it is under pressure from the natural forces
that surround and trap it. If a hole (well) is drilled into the reservoir, an opening is
provided at a much lower pressure through which the reservoir fluids can escape. The
driving force which causes these fluids to move out of the reservoir and into the wellbore
comes from the compression of the fluids that are stored in the reservoir. The actual
energy that causes a well to produce oil comes from the pressure difference between the
reservoir and the production facilities at the surface. The production system can be
divided into several components: the reservoir, the near well bore zone, the well and the
production gathering system. A simple producing system is illustrated in Figure 2.1.

Figure 2.1: Simplified hydrocarbon production system

The hydrocarbon fluid flows from the reservoir into the well, up the tubing, along the
horizontal flow line and into the oil storage tank. During this process the fluid’s pressure
is reduced from the reservoir pressure to atmosphere pressure in a series of pressure loss
processes (Figure 2.2):
(1) across the reservoir
(2) across the completion (perforation/gravel pack etc.)
(3) across the tubing and any restrictions
(4) across the sub surface safety valve
(5) across the surface choke

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 20


Advanced Production Engineering 2019/2020 Academic Year

(6) across flowline

These pressure losses can be grouped into three main components:

(7) summarises the total pressure losses in the reservoir and completion
(8) summarises the total pressure losses in the tubing
(9) summarises the total pressure losses at the surface

Figure 2.2: Pressure losses during production

where PR = Reservoir Pressure


Pwfs = Flowing sand face Pressure
Pwf = Flowing Bottom Hole Pressure
PUR = Upstream Restriction Pressure
PDR = Downstream Restriction Pressure
PUSV = Upstream Safety Valve Pressure
PDSV = Downstream Safety Valve Pressure
PWH = Well Head Pressure
PDSC = Downstream surface Choke Pressure
Psep = Separator Pressure
The magnitude of these individual pressure losses depend on the reservoir properties and
pressures; fluid being produced and the well design. Production engineers need to
understand the interplay of these various factors so as to design completions which
maximise profitability from the oil or gas production. There are no standard “rules of

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 21


Advanced Production Engineering 2019/2020 Academic Year

thumb” which can be used. Figure 2.3 schematically represents the pressure distribution
across the production system shown in Figure 2.2. It identifies the most significant
components, flowline, tubing and the reservoir and completion where pressure losses
occur.

Figure 2.3: Pressure across production system

2.2 NATURAL EQUILIBRIUM FLOWRATE

Pressure, temperature and compositional changes are all important and interrelated but
the pressure losses in the system are of particular interest. The driving force that moves
fluids from the reservoir along the production system is the energy available that is stored
in the form of compressed fluids in the reservoir. As the fluids move along the system
components, pressure drops occur. The pressure in the direction of flow continuously
decreases from the reservoir pressure to the final downstream pressure value at the
separator. The pressure losses (ΔP) occurs due to friction, gravity and acceleration.
Usually all the 3 components of the pressure drop occurs, but the relative importance
varies according to the section of the production system.
Gravitational losses (ΔPgrav) are due to elevation differences and are usually the
dominant pressure loss component in vertical sections such as tubings and risers.
Friction losses (ΔPfric) are due to the shear between the fluid and a solid wall and is
usually the dominant pressure loss component in the porous media and in undersized
flowlines.
Acceleration losses (ΔPacc) are due to velocity changes and are usually present in the
flow of compressible fluids at low pressure, at cross section area changes or during
transient conditions.

ΔP = ΔPgrav + ΔPfric + ΔPacc (2.1)

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 22


Advanced Production Engineering 2019/2020 Academic Year

Usually as the flowrate increases, the pressure drop also increases and the pressure at any
location in the system usually decreases (Figure 2.4a). The summation of all pressure
drops in the production system must be equal to the difference between the upstream and
downstream pressures (Figure 2.4b).
(a)

(b)

Figure 2.4: Pressure losses during production


The flowrate value at which a specific well fluids are lifted to the surface using only the
energy stored in the compressed reservoir fluids is called “natural equilibrium
flowrate”

It is the solution to the following equation (Figure 2.4b):

Pr − Ps =  Pc (q ) (2.2)

The production separator (Figure 2.5) is a large vessel that allows for the separation of
the liquid and gas phase. The separation must be made at a constant pressure and with a
constant liquid level in the separator. The separator has a float that senses the liquid level

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 23


Advanced Production Engineering 2019/2020 Academic Year

and control the liquid outlet. The pressure inside the separator must also be kept constant
and is controlled by a gas outlet valve that is actuated by a control that senses the
separation pressure. If the pressure decreases, the valve closes and if the pressure
increases the valve opens. This manner, the first separator in the separation train can be
considered as a boundary of constant pressure for the production system.

Figure 2.5: Schematics of production separator

The constant separator pressure will control the flowrate that the well is producing
(Figure 2.6).

Figure 2.6: Separator pressures and production flowrates

It is evident from Figure 2.6 that the lower the separator pressure, the bigger the driving
force to move the fluids. As a consequence an inexperienced production engineer is
tempted to lower separator pressure in an attempt to increase production. This can be
done but you must know what you are doing and the consequences of changing
separation conditions. The fluids produced from a well are always a multiphase mixture.
Even the case of production from dead oil reservoirs or dry gas reservoirs may be
associated with some water production. Therefore it is always necessary to have a system
that enable us to separate each individual phase so that they can finally be treated,
measured, transported, stored for future use and or discarded.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 24


Advanced Production Engineering 2019/2020 Academic Year

The liquid-liquid separation is affected by both pressure and temperature. Specifically for
the separation of water and oil, the temperature may play an important role in the case of
emulsions. Therefore, the arriving temperature of the fluids at the separator may
influence the residence time, amount of heat and chemicals required to break an emulsion
and therefore influence the production costs. The temperature that the fluids arrive at the
separation station depends on the fluid properties, flowrates, artificial lift method used,
well and flow line heat transfer properties, external temperature, etc. It may be an
important factor in designing production systems in artic and deep water environments.
In the liquid-liquid separation, the oil and water phases are immiscible and therefore
there is no mass transfer between the two liquid phases.

The gas-liquid separation is different. The hydrocarbon mixture is a complex mixture of


several hydrocarbon substances. At a certain pressure and temperature each hydrocarbon
substance is going to be present in different quantities in the gas and liquid phases.
Usually the gas phase is richer in low molecular weight hydrocarbons and poorer in
higher molecular weight hydrocarbons. On the other hand the liquid phase is richer in
higher molecular weight hydrocarbons and poorer in lower molecular weight
hydrocarbons. Of particular interest is the intermediate molecular weight hydrocarbons
(ethane, propane, butane, pentane, etc…). Depending on the pressure those components
can be more present in the gas or liquid phase and this is an important economic factor.

If more intermediate molecular weight hydrocarbons are present in the liquid phase than
in the gas phase we have the following advantages:

• More hydrocarbons produced as liquid, therefore we increase the mass of liquid


hydrocarbons that can be sold.
• Higher API for the produced liquid due to a higher presence of the intermediate
fractions therefore the liquid has a higher selling price.
• Leaner gas being produced since the intermediate hydrocarbons are in the liquid
tank, therefore the produced gas will not have condensation problems as it is
transported, stored and used as fuel, injected or gas lift gas.

If the separation pressure is too low, then a lot of the higher molecular weight
intermediate hydrocarbons (butane, pentane, etc…) go to the gas phase. There will
therefore be smaller volumes of liquid of lower API to be sold as liquids. The gas stream is
richer in intermediate components that will condense if pressure is further increased or
temperature is lowered. The intermediate hydrocarbons in the gas phase will be sold as
gas. They can be recovered later as hydrocarbon liquids if the process is economically
attractive.

If the separation pressure is too high, then a lot of the lower molecular weight
hydrocarbons (ethane and propane) will be in the liquid phase but we will lose most of
the methane to the gas phase. This in principle is good but later on, in the storage tank,
there is going to be a second separation at a very low pressure and since now we have
almost no methane in the liquid (most of it was sent in the gas phase), the only way we

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 25


Advanced Production Engineering 2019/2020 Academic Year

can increase the partial pressure of the hydrocarbons in the gas phase is by vapourizing
ethane, methane, propane, butane, pentane, etc… Therefore we will lose the intermediate
molecular weight hydrocarbons in the tank which is worse since it is not sold as liquid
and cannot be recovered from the gas stream and just add greenhouse gases to the
atmosphere.

Therefore, the separation of liquid and gas needs to go through a process called
stabilization. In this process, the separation is done at several stages. Usually we have
several separator in series. The liquid stream flow from one separator to the next one at a
lower pressure. The operating pressure of each separator is selected in order to stabilize
the intermediate hydrocarbons in the liquid phase so that they are not lost in the tank and
do not contaminate the produced gas from each separator. The separator pressures in the
stabilization process are very carefully determined by the process engineer to maximize
the economics of the recovered liquids. The facilities engineer is also worried about the
problems related to transporting and compressing the produced gas. The separator
pressures are determined in order to:

• Maximize the recovery of hydrocarbon liquids in the tank;


• Stabilize the gas stream making it leaner;
• Enable the gas stream to be produced to a compression station and;
• Reduce the amount of horsepower required to compress the gas.

This manner, the first separator in the separation train can be considered as a boundary of
constant pressure for the production system (Figure 2.5).

The importance of the stabilization process depends on the nature of the fluids being
produced and depends on the amount of intermediate hydrocarbons present in the fluid
– Black oil
– Volatile oil
– Retrograde condensate
– Wet gas
The separator pressure may be changed but this must be done carefully and after
discussing this with the process and facilities engineers.

The wellhead pressure is an intermediate pressure in the system and is the result of the
equilibrium of losses and available pressure. Some of the factors that may increase a
wellhead pressure are:
• Higher separator pressure: If the high wellhead pressure is being caused by a
high separator pressure this is probably a requirement and very little can be done
about it.

• Choke: If the high wellhead pressure is caused by a choke, we need to determine


the reasons why the choke is restricting the flow. Some possible causes are:

✓ Someone forgot it at a smaller opening;

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 26


Advanced Production Engineering 2019/2020 Academic Year

✓ There is a restriction from reservoir engineers on the production flowrate


from this well;
✓ The well requires some friction losses to be stable.

After determining the cause for the current choke opening we can determine if it is
appropriate or not to reduce the wellhead pressure by opening the choke.

• Excessive losses in the flowline: If the high wellhead pressure is not caused by a
high separator pressure or a choke then the only cause is a high pressure loss in
the flowline. This can be caused by:
✓ Wax, scale or other solid deposits;
✓ The line installed is of insufficient diameter for current production
conditions;
✓ Something is blocking the line such as a lost or stuck pig;
✓ Collapse of the line at some point.

As the reservoir is depleted, its pressure (driving force for producing fluids) naturally
declines. Appearance of a gas phase inside the porous media also reduces the
“productivity” of the liquid oil phase increasing the pressure losses inside the porous
media. There are also changes in producing conditions such as WOR, GOR, deposition of
wax or scale, etc. that may increase the pressure losses through tubing and/or flowlines.
As the well ages, there is a reduction on the ability of the reservoir to deliver fluids after
the perforations at a sufficient pressure to overcome the pressure losses through the
production system. Therefore, the well is constantly seeking an equilibrium point
reducing the flowrate to reduce the pressure losses in the system.

The effects of production decline can be minimized by:


– Pressure maintenance (water or gas injection);
– Properly planned production schedule and reservoir management;
– Stimulation techniques;
– Design of tubing and flowline diameters compatible with production conditions;
– Removal of unnecessary pressure losses;
– Injection of products to inhibit scale or wax formation
– Injection of diluents and thermal heating or insulation of tubing and flowlines to
reduce fluid viscosity; etc.

In all those cases the well is still called a “naturally flowing well (free flowing)” since
the equilibrium flowrate will be determined by the ability of the reservoir pressure to
overcome the pressure losses in the system.

2.3 SYSTEM ANALYSIS OF THE PRODUCTION SYSTEM


The use of systems analysis to design a hydrocarbon production system was first
suggested by Gilbert (“Flowing and Gas Lift Performance”, API Drilling and Production
Practices, 1954”). Systems analysis, which has been applied to many types of systems of
interacting components, consists of selecting a point or node within the producing system

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 27


Advanced Production Engineering 2019/2020 Academic Year

(well and surface facilities). Equations for the relationship between flow rate and pressure
drop are then developed for the well components both upstream of the node (inflow) and
downstream (outflow).

The flow rate and pressure at the node can be calculated since:
(i) Flow into the node equals flow out of the node.
(ii) Only one pressure can exist at the node.
Further, at any time, the pressure at the end points of the system {separator (psep) and
reservoir pressure (pr)} are both fixed. Thus:

pr – (Pressure loss upstream component) = pnode (2.3)


psep + (Pressure loss downstream component) = pnode (2.4)

Typical results of such an analysis is shown in Figure 4 where the pressure-rate


relationship has been plotted for both the inflow (Equation 2.3) and outflow (Equation
2.4) at the node. The intersection of these two lines is the (normally unique) operating
point. This defines the pressure and rate at the node. This approach forms the basis of all
hand and computerised flow calculation procedures. It is frequently referred to as “nodal
analysis”. Chapter 3 discusses this into details.

Figure 2.7: Node flow rate and pressure

CHAPTER 3
RESERVOIR DELIVERABILITY

3.1 INTRODUCTION
Reservoir deliverability is defined as the oil or gas production rate achievable from
reservoir at a given bottom-hole pressure. Reservoir deliverability determines types of
completion and artificial lift methods to be used. A thorough knowledge of reservoir
productivity is essential for production engineers.
Reservoir deliverability depends on several factors including the following:

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 28


Advanced Production Engineering 2019/2020 Academic Year

• Reservoir pressure
• Pay zone thickness and permeability
• Reservoir boundary type and distance
• Wellbore radius
• Reservoir fluid properties
• Near-wellbore condition
• Reservoir relative permeabilities

Reservoir deliverability can be mathematically modelled on the basis of flow regimes such
as transient flow, steady state flow, and pseudo–steady state flow. An analytical relation
between bottom-hole pressure and production rate can be formulated for a given flow
regime. The relation is called ‘‘inflow performance relationship’’ (IPR). This chapter
addresses the procedures used for establishing IPR of different types of reservoirs and
well configurations.

3.2 FLOW REGIMES


When a vertical well is open to produce oil at production rate q, it creates a pressure
funnel of radius ra round the wellbore, as illustrated by the dotted line in Figure 3.1a. In
this reservoir model, the h is the reservoir thickness, k is the effective horizontal reservoir
permeability to oil, µo is viscosity of oil, Bo is oil formation volume factor, rw is wellbore
radius, pwf is the flowing bottom hole pressure, and p is the pressure in the reservoir at the
distance r from the wellbore center line. The flow stream lines in the cylindrical region
form a horizontal radial flow pattern as depicted in Figure 3.1b.

Figure 3.1: A sketch of a radial flow reservoir model: (a) lateral view, (b) top view.

3.2.1 TRANSIENT FLOW

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 29


Advanced Production Engineering 2019/2020 Academic Year

Transient flow is defined as a flow regime where/when the radius of pressure wave
propagation from wellbore has not reached any boundaries of the reservoir. During
transient flow, the developing pressure funnel is small relative to the reservoir size.
Therefore, the reservoir acts like an infinitively large reservoir from transient pressure
analysis point of view. Assuming single-phase oil flow in the reservoir, several analytical
solutions have been developed for describing the transient flow behaviour. A constant-
rate solution expressed by Equation 3.1 is frequently used in production engineering:

162.6qBo  o  k 
p wf = pi −  log t + log − 3.23 + 0.87S  (3.1)
 o ct rw
2
kh  

where
pwf = flowing bottom-hole pressure, psia
pi = initial reservoir pressure, psia
q =oil production rate, stb/day
µo = viscosity of oil, cp
k = effective horizontal permeability to oil, md
h = reservoir thickness, ft
t =flow time, hour
φ = porosity, fraction
ct = total compressibility, psi–1
rw = wellbore radius to the sand face, ft
S = skin factor

Because oil production wells are normally operated at constant bottom-hole pressure
because of constant wellhead pressure imposed by constant choke size, a constant
bottom-hole pressure solution is more desirable for well inflow performance analysis.
With an appropriate inner boundary condition arrangement, Earlougher (1977)
developed a constant bottom-hole pressure solution, which is similar to Equation 3.1:

kh( pi − p wf )
q= (3.2)
 k 
162.6 Bo  o  log t + log − 3.23 + 0.87S 
 o ct rw
2

which is used for transient well performance analysis in production engineering. Equation
3.2 indicates that oil rate decreases with flow time. This is because the radius of the
pressure funnel, over which the pressure drawdown (pi – pwf) acts, increases with time,
that is, the overall pressure gradient in the reservoir drops with time. For gas wells, the
transient solution is:

qg =

kh m( pi ) − m( p wf )  (3.3)
 k 
1,638T  log t + log − 3.23 + 0.87S 
  o ct rw
2

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 30


Advanced Production Engineering 2019/2020 Academic Year

where qg is production rate in Mscf/d, T is temperature in oR, and m(p) is real gas pseudo-
pressure defined as:

m( p ) =
2p
 z dp
pb
(3.4)

3.2.2 STEADY-STATE FLOW


Steady-state flow is defined as a flow regime where the pressure at any point in the
reservoir remains constant over time. This flow condition prevails when the pressure
funnel shown in Figure 3.1 has propagated to a constant pressure boundary. The
constant-pressure boundary can be an aquifer or a water injection well. A sketch of the
reservoir model is shown in Figure 3.2, where pe represents the pressure at the constant-
pressure boundary. Assuming single-phase flow, the following theoretical relation can be
derived from Darcy’s law for an oil reservoir under the steady-state flow condition due to
a circular constant pressure boundary at distance re from wellbore:

kh( pe − p wf )
q= (3.5)
 r 
141.2 Bo  o  In e + S 
 rw 

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 31


Advanced Production Engineering 2019/2020 Academic Year

Figure 3.2: A sketch of a reservoir with a constant-pressure boundary.

3.2.3 PSEUDO–STEADY–STATE FLOW


Pseudo–steady-state flow is defined as a flow regime where the pressure at any point in
the reservoir declines at the same constant rate over time. This flow condition prevails
after the pressure funnel shown in Fig. 3.1 has propagated to all no-flow boundaries. A no-
flow boundary can be a sealing fault, pinch-out of pay zone, or boundaries of drainage
areas of production wells. A sketch of the reservoir model is shown in Figure 3.3, where pe
represents the pressure at the no-flow boundary at time t4. Assuming single-phase flow,
the following theoretical relation can be derived from Darcy’s law for an oil reservoir
under pseudo–steady-state flow condition due to a circular no-flow boundary at distance
re from wellbore:
kh( pe − p wf )
q= (3.6)
 re 
141.2 Bo  o  In − 0.5 + S 
 rw 

Figure 3.3: A sketch of a reservoir with no-flow boundaries.

The flow time required for the pressure funnel to reach the circular boundary can be
expressed as:

o ct re2
t pss = 1,200 (3.7)
k

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 32


Advanced Production Engineering 2019/2020 Academic Year

Because the pe in Equation 3.6 is not known at any given time, the following expression
using the average reservoir pressure is more useful:

q=
(
kh p − p wf ) (3.8)
 r 
141.2 Bo  o  In e − 0.75 + S 
 rw 
where p is the average reservoir pressure in psia.

If the no-flow boundaries delineate a drainage area of noncircular shape, the following
equation should be used for analysis of pseudo–steady-state flow:

q=
(
kh p − p wf ) (3.9)
1 4A 
141.2 Bo  o  In + S 
 2 C A rw
2

where A = drainage area, ft2


γ = 1.78 = Euler’s constant
CA = drainage area shape factor, 31.6 for a circular boundary.

The value of the shape factor CA can be found from Figure 3.4.

For a gas well located at the center of a circular drainage area, the pseudo–steady-state
solution is:

qg =

kh m( pi ) − m( p wf )  (3.10)
 r 
1,424T  In e − 0.75 + S + Dq g 
 rw 

where
D = non-Darcy flow coefficient, d/Mscf

3.2.4 HORIZONTAL WELL


The transient flow, steady-state flow, and pseudo–steadystate flow can also exist in
reservoirs penetrated by horizontal wells. Different mathematical models are available
from literature. Joshi (1988) presented the following relationship considering steady-
state flow of oil in the horizontal plane and pseudo–steady-state flow in the vertical plane:

k H h( p e − p wf )
q= (3.11)
  a + a 2 + (L 2 )  I h  I h  
141.2 B  In   + ani In  ani

  L 2  L  w ani + 1)  
r ( I

where

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 33


Advanced Production Engineering 2019/2020 Academic Year

L 0.5
a =   0.5 + 0.25 + (2reh L )  (3-12)
4

 2   
with
43,560A
reh =

and
kH
I ani = (3-13)
kV

kH = the average horizontal permeability, md


kV = vertical permeability, md
reh = radius of drainage area, ft
L = length of horizontal wellbore (L=2<0.9 reh), ft.

Figure 3.4: Shape factors: (a) closed drainage areas with low-aspect ratios. (b) closed
drainage areas with high-aspect ratios (Dietz, 1965).

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 34


Advanced Production Engineering 2019/2020 Academic Year

3.3 INFLOW PERFORMANCE RELATIONSHIP


IPR is used for evaluating reservoir deliverability in production engineering. The IPR
curve is a graphical presentation of the relation between the flowing bottom-hole
pressure and liquid production rate. A typical IPR curve is shown in Figure 3.5. The
magnitude of the slope of the IPR curve is called the ‘‘productivity index’’ (PI or J), that is:

q
J= (3.14)
( pe − p wf )
where J is the productivity index. Apparently J is not a constant in the two-phase flow
region.

Figure 3.5: A typical IPR curve for an oil well.

3.3.1 IPR FOR SINGLE (LIQUID)-PHASE RESERVOIRS


All reservoir inflow models represented by Equations 3.1, 3.3, 3.7, and 3.8 were derived
on the basis of the assumption of single-phase liquid flow. This assumption is valid for
undersaturated oil reservoirs, or reservoir portions where the pressure is above the
bubble-point pressure. These equations define the productivity index (J) for flowing
bottom-hole pressures above the bubble-point pressure as follows.

For radial transient flow around a vertical well:


q kh
J= = (3.15)
( pi − p wf )  k 
162.6 Bo  o  log t + log − 3 . 23 + 0 . 87 S 

  c r
o t w
2

For radial steady-state flow around a vertical well:

q kh
J=
( pe − p wf ) =  r 
141.2 Bo  o  In e + S 
(3.16)

 rw 

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 35


Advanced Production Engineering 2019/2020 Academic Year

For pseudo–steady-state flow around a vertical well:


q kh
J= = (3.17)
(
p − p wf ) 1
141.2 Bo  o  In
4A 
+ S 
 2 C A rw
2

and for steady-state flow around a horizontal well:

q kH h
J=
( pe − p wf ) =   a + a 2 + (L 2)  I h  I h  
(3.18)
141.2 B  In   + ani In  ani

  L 2  L  w ani + 1)  
r ( I

Since the productivity index J above the bubble-point pressure is independent of


production rate, the IPR curve for a single (liquid)-phase reservoir is simply a straight line
drawn from the reservoir pressure to the bubble-point pressure. If the bubble-point
pressure is 0 psig, the absolute open flow (AOF) is the productivity index J times the
reservoir pressure.

Example 3.1
Construct IPR of a vertical well in an oil reservoir. Consider (1) transient flow at 1 month,
(2) steady-state flow, and (3) pseudo–steady-state flow. The following data are given:

Φ = 0.19 k = 8.2 md h = 53 ft pe or p = 5,651 psia pb = 50 psia


Bo = 1.1 µo =1.7 cp A = 640 acres rw =0.328 ft S=0
ct =0.0000129 psi–1

Solution
1. For transient flow, calculated points are:

kh
J=
 k 
162.6 Bo  o  log t + log − 3.23 + 0.87S 
  o ct rw
2

J=
(8.2)(53)
 
162.6(1.1)(1.7 ) log(30  24) + log
8 .2
− 3.23
 (0.19)(1.7 )(0.0000129)(0.328) 2

J = 0.2075STB/day/psi

Therefore qo = 0.2075 (pi – pwf)

The calculated points are:

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 36


Advanced Production Engineering 2019/2020 Academic Year

pwf (psi) qo (stb/day)


5,651 0
50 1,162

Transient IPR curve is plotted in Figure 3.6.

Figure 3.6: Transient IPR curve for Example 3.1.

2. For steady state flow:


kh
J=
 r 
141.2 Bo  o  In e + S 
 rw 

J=
(8.2)(53) = 0.1806 STB/day/psi
 2,980 
141.2(1.1)(1.7 )In 
 0.328 

Therefore qo = 0.1806 (pe – pwf)

The calculated points are:

pwf (psi) qo (stb/day)


5,651 0
50 1,011

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 37


Advanced Production Engineering 2019/2020 Academic Year

Figure 3.7: Steady-state IPR curve for Example 3.1.

3. For pseudosteady state flow

kh
J=
 r 
141.2 Bo  o  In e − 0.75 + S 
 rw 

J=
(8.2)(53) = 0.1968 STB/day-psi
 2,980 
141.2(1.1)(1.7 ) In − 0.75
 0.328 

Therefore qo = 0.1968 ( p – pwf)

Calculated points are:

pwf (psi) qo (stb/day)


5,651 0
50 1,102

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 38


Advanced Production Engineering 2019/2020 Academic Year

Figure 3.8: Pseudo–steady-state IPR curve for Example 3.1.

3.3.2 IPR FOR RESERVOIRS USING EMPIRICAL EQUATIONS

VOGEL’S METHOD

Vogel’s equation is still widely used in the industry. It is written as:

qo   p wf  p 
2

= 1 − 0.2  − 0.8 wf   (3.19)
(qo )max 
  pr


 p
 r

 

where
qo=oil rate at pwf
(qo)max=maximum oil flow rate at zero wellbore pressure, i.e., AOF
p r =current average reservoir pressure, psig
pwf = wellbore pressure, psig

  qo  
p wf = 0.125 p r  81 − 80  − 1 (3.20)
 (q o )max
  
  

Vogel’s methodology can be used to predict the IPR curve for the following two types of
reservoirs:
• Saturated oil reservoirs p r ≤ pb
• Undersaturated oil reservoirs p r > pb

Saturated Oil Reservoirs


When the reservoir pressure equals the bubble-point pressure, the oil reservoir is
referred to as a saturated oil reservoir. The computational procedure of applying Vogel’s
method in a saturated oil reservoir to generate the IPR curve for a well with a stabilized
flow data point, i.e., a recorded qo value at pwf, is summarized below:

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 39


Advanced Production Engineering 2019/2020 Academic Year

Step 1. Using the stabilized flow data, i.e., qo and pwf, calculate: (qo)max from Equation 3.19
or
qo   p wf  p 
2

= 1 − 0.2  − 0.8 wf  
(qo )max 
  pr


 p
 r

 

Step 2. Construct the IPR curve by assuming various values for pwf and calculating the
corresponding qo from:

  p wf  p 
2

q o = (q o )max 1 − 0.2  − 0.8 wf
  p



  pr   r  
 

Example 3.2
A well is producing from a saturated reservoir with an average reservoir pressure of
2,500 psig. Stabilized production test data indicated that the stabilized rate and wellbore
pressure are 350 STB/day and 2,000 psig, respectively. Calculate:
(i) Oil flow rate at pwf =1850 psig
(ii) Calculate oil flow rate assuming constant J
(iii) Construct the IPR by using Vogel’s method and the constant productivity index
approach.

Solution
(i) Step 1. Calculate (qo)max:

 p  p 
2

(qo )max = qo 1 − 0.2 wf
 p
 − 0.8 wf
  p



  r   r  
 

  2,000   2,000  
2

(qo )max = 350 1 − 0.2  − 0.8  


  2,500   2,500  

= 1067.1STB/day

Step 2. Calculate qo at pwf = 1850 psig by using Vogel’s equation

  p wf  p 
2

q o = (q o )max 1 − 0.2  − 0.8 wf
  p



  pr   r  
 
  1,850   1,850  
2

qo = 1067.11 − 0.2  − 0.8  


  2,500   2,500  

= 441.7 STB/day

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 40


Advanced Production Engineering 2019/2020 Academic Year

(ii) Calculating oil flow rate by using the constant J approach

Step 1. Apply Equation 3.14 to determine J

qo 350
J= = = 0.7 STB/day/psi
(p r − p wf ) 2,500 − 2,000

Step 2. Calculate qo

( )
q o = J p r − p wf = 0.7(2,500 − 1,850) = 455 STB/day

(iii) Generating the IPR by using the constant J approach and Vogel’s method:
Assume several values for pwf and calculate the corresponding qo

Undersaturated Oil Reservoirs


Beggs (1991) pointed out that in applying Vogel’s method for undersaturated reservoirs,
there are two possible outcomes to the recorded stabilized flow test data that must be
considered, as shown schematically in Figure 3.9.

• The recorded stabilized bottom-hole flowing pressure is greater than or equal to


the bubble-point pressure, i.e. pwf ≥pb

• The recorded stabilized bottom-hole flowing pressure is less than the bubble-point
pressure pwf < pb

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 41


Advanced Production Engineering 2019/2020 Academic Year

Figure 3.9: Stabilized flow test data.

Case 1. The Value of the Recorded Stabilized pwf ≥pb


Beggs outlined the following procedure for determining the IPR when the stabilized
bottom-hole pressure is greater than or equal to the bubble point pressure (Figure 3.9):

Step 1. Using the stabilized test data point (qo and pwf) calculate the productivity index J:
qo
J=
p r − p wf

Step 2. Calculate the oil flow rate at the bubble-point pressure:

(
qob = J p r − pb ) (3.21)

where qob is the oil flow rate at pb

Step 3. Generate the IPR values below the bubble-point pressure by assuming different
values of pwf < pb and calculating the corresponding oil flow rates by applying the
following relationship:

Jp  
2
 p wf  p 
q o = q ob + b 1 − 0.2  − 0.8 wf
  p


 (3.22)
1.8   pr   r  
 

The maximum oil flow rate [(qo)max or AOF] occurs when the bottomhole flowing pressure
is zero, i.e. pwf =0, which can be determined from the above expression as:

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 42


Advanced Production Engineering 2019/2020 Academic Year

Jpb
(qo )max = qob +
1.8

It should be pointed out that when pwf ≥pb, the IPR is linear and is described by: qo=
J( p r − pwf) as already shown in Section 3.3.1.

Example 3.3
Construct IPR of a vertical well in a saturated oil reservoir using Vogel’s equation,
assuming pseudosteady flow. The following data are given:
Φ = 0.19 k = 8.2 md h = 53 ft p = 5,651 psia pb = 5,651 psia
Bo = 1.1 µo = 1.7 cp rw = 0.328 ft S=0 A = 640 acres ct =
0.0000129 psi–1

Solution
kh
J=
 r 
141.2 Bo  o  In e − 0.75 + S 
 rw 

J=
(8.2)(53) = 0.1968 STB/day/psi
 2,980 
141.2(1.1)(1.7 ) In − 0.75
 0.328 

( )
qob = J p r − pb = 0.1968(0) = 0

Jpb (0.1968)(5,651)
= = 618STB/day
1.8 1.8

Generate the IPR values below the bubble-point pressure by assuming different values of
pwf < pb and calculating the corresponding oil flow rates by applying:
Jpb   p wf  
2
 p wf 
q o = q ob + 1 − 0.2    
1.8   p  − 0.8 p  
  r   r  
  p wf   p wf  
2

q = 0 + 6181 − 0.2  − 0.8  


  5, 651   5, 651  

The calculated points are:

pwf (psi) qo (stb/day)

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 43


Advanced Production Engineering 2019/2020 Academic Year

5,651 0
5,000 122
4,500 206
4,000 283
3,500 352
3,000 413
2,500 466
2,000 512
1,500 550
1,000 580
500 603
0 618

The IPR curve is plotted in Figure 3.10.

Figure 3.10: IPR curve for Example 3.3.

Example 3.4
Construct IPR of a vertical well in an undersaturated oil reservoir using the generalized
Vogel equation and assuming pseudosteady flow. The following data are given:
Φ = 0.19 k = 8.2 md h = 53 ft p = 5,651 psia pb = 3,000 psia
Bo = 1.1 µo = 1.7 cp rw = 0.328 ft S=0 A = 640 acres
ct = 0.0000129 psi–1

Solution
kh
J=
 r 
141.2 Bo  o  In e − 0.75 + S 
 rw 

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 44


Advanced Production Engineering 2019/2020 Academic Year

J=
(8.2)(53) = 0.1968 STB/day/psi
 2,980 
141.2(1.1)(1.7 ) In − 0.75
 0.328 

(
qob = J p − pb )
= 0.1968(5,651− 3,000)
= 522 STB/day

Jpb (0.1968)(3,000)
= = 328 STB/day
1.8 1.8

Generate the IPR values below the bubble-point pressure by assuming different values of
pwf < pb and calculating the corresponding oil flow rates by applying:

Jp  
2
 p wf  p 
q o = q ob + b 1 − 0.2  − 0.8 wf  
1 .8   pb   pb  

  p wf   p wf  
2

q = 522 + (328)1 − 0.2  − 0.8  


  3,000   3,000  

The calculated points are:


pwf (psi) qo (stb/day)
5,651 0
3,000 522
2,826 555
2,260 651
1,695 729
1,130 788
565 828
0 850

The IPR curve is plotted in Figure 3.11.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 45


Advanced Production Engineering 2019/2020 Academic Year

Figure 3.11: IPR curve for Example 3.4.

Example 3.5
An oil well is producing from an undersaturated reservoir that is characterized by a
bubble-point pressure of 2,130 psig. The current average reservoir pressure is 3,000 psig.
Available flow test data show that the well produced 250 STB/day at a stabilized pwf of
2,500 psig. Construct the IPR data.

Solution
The problem indicates pwf>pb, therefore, the Case 1 procedure for undersaturated
reservoirs as outlined previously must be used.

Step 1. Calculate J using the flow test data.


qo 250
J= = = 0.5 STB/day/psi
p r − p wf 3,000 − 2,500

Step 2. Calculate the oil flow rate at the bubble-point pressure.

( )
qob = J p r − pb = 0.5(3,000 − 2,130) = 435STB/day

Jpb (0.5)(2,130)
= = 591.7 STB/day
1.8 1.8

Generate the IPR values below the bubble-point pressure by assuming different values of
pwf < pb and calculating the corresponding oil flow rates by applying:

Jp  
2
 p wf  p 
q o = q ob + b 1 − 0.2  − 0.8 wf  
1 .8   pb   pb  

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 46


Advanced Production Engineering 2019/2020 Academic Year

  p wf   p wf  
2

q = 435 + (591.7 )1 − 0.2  − 0.8   (*)


  2,130   2,130  

Calculate IPR values above bubble-point pressure (pwf > pb), the IPR is linear and is
described by: qo= J( p r − pwf) (**)
pwf Eqn. qo (stb/day)
3000 (**) 0
2800 (**) 100
2600 (**) 200
2130 (**) 435
1500 (*) 709
1000 (*) 868
500 (*) 973
0 (*) 1027

Figure 3.12: IPR curve for Example 3.5.

Case 2. The Value of the Recorded Stabilized pwf < pb


When the recorded pwf from the stabilized flow test is below the bubble-point pressure, as
shown in Figure 3.9, the following procedure for generating the IPR data is proposed:

Step 1. Using the stabilized well flow test data and combining Equation 3.21 with 3.22,
solve for the productivity index J to give:

qo
J= (3.23)
p  
2

( )  p wf
p r − pb + b 1 − 0.2
1.8 
 p
 − 0.8 wf

 
  pb   pb  

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 47


Advanced Production Engineering 2019/2020 Academic Year

Step 2. Calculate qob by using:

(
qob = J p r − pb )
Step 3. Generate the IPR for pwf ≥pb by assuming several values for pwf above the bubble
point pressure and calculating the corresponding qo from:

(
q o = J p r − p wf )
Step 4. Calculate qo at various values of pwf below pb, or:
Jpb   p wf  
2
 p wf 
q o = q ob + 1 − 0.2  − 0.8  
1 .8   pb   pb  
 

Example 3.6
The well described in Example 3.5 was retested and the following results obtained:
pwf =1700 psig, qo= 630.7 STB/day
Generate the IPR data using the new test data.

Solution
p r = 3,000 psig pb =2,130 psig pwf =1700 psig, qo= 630.7 STB/day

Notice that the stabilized pwf is less than pb.

Step 1. Solve for J by applying Equation 3.23.


qo
J=
p  
2

( )  p wf 
p r − pb + b 1 − 0.2
1.8 
p
 − 0.8 wf

 
  pb   pb  

630.7
J= = 0.5 STB/day/psi
 2

(3,000 − 2,130) + 2,130 1 − 0.2 1,700  − 0.8 1,700  
1.8   2,130   2,130  

Step 2. Calculate qob by using:

( )
qob = J p r − pb = 0.5(3,000 − 2,130) = 435 STB/day

Step 3. Generate the IPR for pwf ≥pb by assuming several values for pwf above the bubble
point pressure and calculating the corresponding qo from:

(
q o = J p r − p wf ) (**)

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 48


Advanced Production Engineering 2019/2020 Academic Year

Step 4. Calculate qo at various values of pwf below pb, or:


Jpb   p wf  
2
 p wf 
q o = q ob + 1 − 0.2  − 0.8   (*)
1 .8   pb   pb  
 
pwf Eqn. qo (stb/day)
3000 (**) 0
2800 (**) 100
2600 (**) 200
2130 (**) 435
1500 (*) 709
1000 (*) 867
500 (*) 973
0 (*) 1027
Trial
Construct IPR of two wells in an undersaturated oil reservoir using the generalized Vogel
equation. The following data are given:
Reservoir pressure: p r = 5,000 psia
Bubble point pressure: pb = 3,000 psia
Tested flowing bottom-hole pressure in Well A: pwfa = 4,000 psia
Tested production rate from Well A: qa = 300 stb/day
Tested flowing bottom-hole pressure in Well B: pwfb = 2,000 psia
Tested production rate from Well B: qb= 900 stb/day

Solution
Well A:
Step 1. Calculate J using the flow test data.
qo 300
J= = = 0.3 STB/day/psi
p r − p wf 1 5,000 − 4,000

Step 2. Calculate the oil flow rate at the bubble-point pressure.

( )
qob = J p r − pb = 0.3(5,000 − 3,000) = 600 STB/day

Jpb (0.3)(3,000)
= = 500 STB/day
1.8 1.8

Generate the IPR values below the bubble-point pressure by assuming different values of
pwf < pb and calculating the corresponding oil flow rates by applying:

Jp  
2
 p wf  p 
q o = q ob + b 1 − 0.2  − 0.8 wf  
1 .8   pb   pb  

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 49


Advanced Production Engineering 2019/2020 Academic Year

The calculated points are:


pwf qo (stb/day)
5000 0
3000 600
2500 739
2000 856
1500 950
1000 1,022
500 1,072
0 1,100

Figure 3.13: IPR curves for Well A.


Well B:
Step 1. Solve for J by applying Equation 3.23.
qo
J=
p  
2

( )  p wf 
p r − pb + b 1 − 0.2
1.8 
p
 − 0.8 wf

 
  pb   pb  

900
J= = 0.3156 STB/day/psi
3,000   2,000  
2
 2,000 
(5,000 − 3,000) + 1 − 0.2  − 0.8  
1.8   3,000   3,000  

Step 2. Calculate qob by using:

( )
qob = J p r − pb = 0.32(5,000 − 3,000) = 631 STB/day

Generate the IPR values below the bubble-point pressure by assuming different values of
pwf < pb and calculating the corresponding oil flow rates by applying:

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 50


Advanced Production Engineering 2019/2020 Academic Year

Jp  
2
 p wf  p 
q o = q ob + b 1 − 0.2  − 0.8 wf  
1 .8   pb   pb  

The calculated points are:
pwf qo (stb/day)
5000 0
3000 631
2500 777
2000 900
1500 999
1000 1,075
500 1,128
0 1,157

Figure 3.14: IPR curves for Well B.

FETKOVICH’S METHOD
Muskat and Evinger (1942) attempted to account for the observed nonlinear flow
behaviour (i.e., IPR) of wells by calculating a theoretical productivity index from the
pseudosteady-state flow equation. They expressed Darcy’s equation as:

(3.24
)

where the pressure function f(p) is defined by:

(3.25
)

where kro = oil relative permeability


k = absolute permeability, md

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 51


Advanced Production Engineering 2019/2020 Academic Year

Bo = oil formation volume factor


μo = oil viscosity, cp

Fetkovich (1973) suggests that the pressure function f(p) can basically fall into one of the
following two regions:

Region 1: Undersaturated Region


The pressure function f(p) falls into this region if p > pb. Since oil relative permeability in
this region equals unity (i.e., kro = 1), then:

(3.26
)

Fetkovich observed that the variation in f(p) is only slight and the pressure function is
considered constant as shown in Figure 3.15.

Region 2: Saturated Region Region: Undersaturated Region

Figure 3.15: Pressure function concept.

Region 2: Saturated Region


In the saturated region where p < pb, Fetkovich shows that the (kro/μoBo) changes linearly
with pressure and that the straight line passes through the origin. This linear is shown
schematically in Figure 3.15 and can be expressed mathematically as:

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 52


Advanced Production Engineering 2019/2020 Academic Year

(3.27
)

where μo and Bo are evaluated at the bubble-point pressure. In the application of the
straight-line pressure function, there are three cases that must be considered:
• p r and pwf > pb
• p r and pwf < pb
• p r > pb and pwf < pb

Case 1: p r and pwf > pb


This is the case of a well producing from an undersaturated oil reservoir where both pwf
and p r are greater than the bubble-point pressure. The pressure function f(p) in this case
is described by Equation 3.26. Substituting Equation 3.26 into Equation 3.24 gives:

Since (1/µoBo) is constant, then:

(3.28
)

or
(
q o = J p r − p wf ) (3.29
)

The productivity index is defined in terms of the reservoir parameters as:

(3.30
)

(
where Bo and μo are evaluated at p r + p wf 2 )
Example 3.7
A well is producing from an undersaturated oil reservoir that exists at an average
reservoir pressure of 3,000 psi. The bubble-point pressure is recorded as 1,500 psi at
150°F. The following additional data are available:
Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 53
Advanced Production Engineering 2019/2020 Academic Year

• stabilized flow rate = 280 STB/day


• stabilized wellbore pressure = 2200 psi
• h = 20′ rw = 0.3′ re = 660′ s = −0.5 k =65 md
• μo at 2600 psi = 2.4 cp
• Bo at 2600 psi =1.4 bbl/STB

Calculate the productivity index by using both the reservoir properties (i.e., Equation
3.30) and flow test data (i.e., Equation 3.29).

Solution
• From reservoir properties:

• From production data:

(
q o = J p r − p wf )

Results show a reasonable match between the two approaches. It should be noted,
however, that there are several uncertainties in the values of the parameters used in
Equation 3.30 to determine the productivity index. For example, changes in the skin factor
k or drainage area would change the calculated value of J.

Case 2: p r and pwf < pb


When the reservoir pressure p r and bottom-hole flowing pressure pwf are both below the
bubble-point pressure pb, the pressure function f(p) is represented by the straight-line
relationship as expressed by Equation 3.27. Combining Equation 3.27 with Equation 3-24
gives:

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 54


Advanced Production Engineering 2019/2020 Academic Year

 1  1 
Since the term    is constant, then:
  o Bo  pb 

Integrating gives:

Qo =
0.00708kh
  re 
 1

  2 pb
(
 2
 p r − p wf
2
) (3.31)

( o Bo ) pb  In  − 0.75 + s 
  rw  

Introducing the productivity index into the above equation gives:

(3.32
)

 J 
The term   is commonly referred to as the performance coefficient C, or:
 2 pb 

Qo can therefore be:

(3.33
)
To account for the possibility of non-Darcy flow (turbulent flow) in oil wells, Fetkovich
introduced the exponent n in Equation 3.34 to yield:

(3.34
)

The value of n ranges from 1.0 for a complete laminar flow to 0.5 for highly turbulent flow.

There are two unknowns in Equation 3.34: the performance coefficient C and the
exponent n. At least two tests are required to evaluate these two parameters, assuming p r

is known.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 55


Advanced Production Engineering 2019/2020 Academic Year

By taking the log of both sides of Equation 3.34 and solving for log (p r
2 2
)
− p wf , the
expression can be written as:

log Qo = log C + n log p r − p wf


2
( 2
)

(
A plot of p r − p wf
2
2
)
versus Qo on log-log scales will result in a straight line having a slope
of 1/n and an intercept of C at p r − pwf = 1. The value of C can also be calculated using any
2 2

point on the linear plot once n has been determined to give:

Qo
C=
(p 2
r − p wf
2
)
n

Once the values of C and n are determined from test data, Equation 3.34 can be used to
generate a complete IPR.

To construct the future IPR when the average reservoir pressure declines to p r ( ), f

Fetkovich assumes that the performance coefficient C is a linear function of the average
reservoir pressure and, therefore, the value of C can be adjusted as:

(C ) f
 pr
= (C ) p 
( ) f


( )
 p r p 

(3.35)

where the subscripts f and p represent the future and present conditions.

Example 3.8
A four-point stabilized flow test was conducted on a well producing from a saturated
reservoir that exists at an average pressure of 3,600 psi.
Qo (stb/day) pwf, psi
263 3170
383 2890
497 2440
640 2150

i. Construct a complete IPR by using Fetkovich’s method.


ii. Construct the IPR when the reservoir pressure declines to 2,000 psi.

Solution
Part A
Step 1. Construct the following table:

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 56


Advanced Production Engineering 2019/2020 Academic Year

(
Step 2. Plot p r − p wf
2 2
)verses Qo on log-log paper as shown in Figure 3.16 and determine
the exponent n, or:

Qo

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 57


Advanced Production Engineering 2019/2020 Academic Year

Figure 3.16: Flow-after-flow data for Example 3.8 (After Beggs, D., “Production
Optimization Using Nodal Analysis).

Step 3. Solve for the performance coefficient C:

C = 0.00079

Step 4. Generate the IPR by assuming various values for pwf and calculating the
corresponding flow rate from:

(
Qo = 0.00079 3,6002 − p wf
2
)0.854

pwf qo (stb/day)
3600 0
3000 340
2500 534
2000 684
1500 796
1000 875
500 922
0 937

The IPR curve is shown in Figure 3.17. Notice that the AOF, i.e.,(Qo)max, is 937 STB/day.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 58


Advanced Production Engineering 2019/2020 Academic Year

Figure 3.17: IPR using Fetkovich method.

Part B
Step 1. Calculate future C by applying Equation 3.35:
( )
 p 
(C ) f = (C ) p  r f 
( )
 p r p 
 2,000 
(C ) f = 0.000079  = 0.000439
 3,600 

Step 2.Construct the new IPR curve at 2,000 psi by using the new calculated C and
applying the inflow equation.

(
Qo = 0.000439 2,0002 − p wf
2
)
0.854

pwf qo (stb/day)
2000 0
1500 94
1000 150
500 181
0 191

Both the present time and future IPRs are plotted in Figure 3.18.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 59


Advanced Production Engineering 2019/2020 Academic Year

Figure 3.18: Present and Future IPRs.

Case 3: p r > pb and pwf < pb


Figure 3.19 shows a schematic illustration of Case 3 in which it is assumed that pwf < pb
and p r > pb. The integral in Equation 3.24 can be expanded and written as:

Figure 3.19: (kro/μoBo) vs. pressure for Case 3.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 60


Advanced Production Engineering 2019/2020 Academic Year

Substituting Equations 3.26 and 3.27 into the above expression gives:

where μo and Bo are evaluated at the bubble-point pressure pb.

Arranging the above expression gives:

Integrating and introducing the productivity index J into the above relationship gives:

or

(
Qo = J  p r − pb + )
1
( 2 
pb2 − p wf) (3.36)
 2 pb 

Example 3.9
The following reservoir and flow test data are available on an oil well:
• Pressure data: p r = 4000 psi pb = 3200 psi
• Flow test data: pwf = 3600 psi Qo= 280 STB/day

Generate the IPR data of the well.

Solution
Step 1. Calculate the productivity index from the flow test data.
qo 280
J= = = 0.7 STB/day/psi
(
p r − p wf )4,000 − 3,600

Step 2. Generate the IPR data by applying Equation 3.29 when the assumed p wf > pb and
using Equation 3.36 when pwf < pb.

(
q o = J p r − p wf ) (Eqn. 3.29)

( )
Qo = J  p r − pb +
1
( 2 
)
pb2 − p wf  (Eqn. 3.36)
 2 pb 
pwf Eqn. qo (stb/day)
4000 (3.29) 0
3800 (3.29) 140
3600 (3.29) 280
3200 (3.29) 560
3000 (3.36) 696

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 61


Advanced Production Engineering 2019/2020 Academic Year

2600 (3.36) 941


2200 (3.36) 1151
2000 (3.36) 1243
1000 (3.36) 1571
500 (3.36) 1653
0 (3.36) 1680

Results of the calculations are shown graphically in Figure 3.20.

Figure 3.20: IPR for Example 3.9.

3.3.3 EFFECT OF SELECTED PARAMETERS ON IPR


A) Tubing size and liquid loading
The well production will normally increase as the tubing size increases. However, at a
certain point the upward fluid flow velocity will decrease so much (due to the tubing
diameter increase) that it is no longer sufficient to efficiently lift the liquid to the surface
i.e. slip phenomena commence and liquid holdup (or liquid loading) begins (Figure 3.21).

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 62


Advanced Production Engineering 2019/2020 Academic Year

Figure 3.21: Liquid loading analysis

B) Effect of Water Cut and Depletion


An increasing water cut reduces the gas liquid ratio as well as increasing the hydrostatic
head between the reservoir and the surface. This is illustrated in Figure 3.22 for a slightly
over pressured reservoir.

Figire 3.22: Effect of wate cut on production.

Reservoir simulation can be used to predict the reservoir pressure depletion with time
along with any increase in water cut. Such a simulation is illustrated in Table below:

The effect of pressure depletion on the production rate is summarised in Figure 3.23a and
the two are combined in Figure 3.23b. The production rate at time t3 is only 25% of the

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 63


Advanced Production Engineering 2019/2020 Academic Year

initial production, while a small further reduction in reservoir pressure or increase in


water cut beyond 50% will cause the well to cease production altogether.

Figire 3.23: (a) Effect of pressure depletion (b) Effect of pressure depletion and water

C) Effect of skin
Figure 3.24 shows the current well inflow (skin = +8) together with its partial well inflow
(skin = +2) and complete well inflow (skin = 0) removal. The carrying out of a hydraulic
fracture (skin = –3) is also illustrated.

Figire 3.24: Effect of skin factor on well performance.


Tubing 2 (with a more vertical outflow profile) is already restricting production with the
impaired (skin = +8), while only minor (probably uneconomic) production gains are
recorded when the skin is removed. The most favourable production (-3) being still less
than that achieved with the larger tubing.

D) Effect of perforations
The high skin discussed previously under Case C could have many causes e.g. formation
damage, partial completion etc. One factor under the control of the production engineer is

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 64


Advanced Production Engineering 2019/2020 Academic Year

the number and type of perforations. Figure 3.25a illustrates an increase in the numbers
of perforations {N1<N2<N3<N4, all of diameter D1} while Figure 3.25b shows that effect of
a restricted number of perforations (N1) can be (partially) compensated for by an increase
in the diameter from D1 to D3. Theoretically increasing the number of perforations is more
beneficial since it improves the inflow from the reservoir as well as decreasing the
(average) frictional pressure drop in the perforation tunnel (see Figure 3.25c for
comparison). The cost of the perforation operation will increase as the number and
diameter of the perforations are increased - an economic optimum will be found when
both factors are varied simultaneously.

(a) (b)

(c)

Figure 3.25: (a) Effect of number of perforations on production rate, (b) Effect of
perforation diameter on production rate (c) Optimisation of perforating schedule.

E) Effect of wellhead pressure


The separator pressure is often the main component in the surface pressure losses. It
exerts a restrictive “back pressure” on the well production which limits the total pressure
drop available for fluid inflow from the reservoir and onward transportation to the
surface. This effect is illustrated in Figure 3.26. Reducing the separator pressure is often
an effective way of increasing the well production.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 65


Advanced Production Engineering 2019/2020 Academic Year

Figure 3.26: Effects of separator pressure on production rate

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 66


Advanced Production Engineering 2019/2020 Academic Year

CHAPTER 4

ARTIFICIAL LIFT

4.1 INTRODUCTION

Production wells are free flowing or lifted. A free flowing oil well has enough downhole
pressure to reach a suitable wellhead production pressure and maintain an acceptable
well-flow. If the formation pressure is too low, and water or gas injection cannot maintain
pressure or is not suitable, then the well must be artificially lifted. Artificial lift consists of
the installation of specific devices that will help the reservoir bottom hole pressure to
overcome the pressure losses in the system downstream of the perforations.

There are some naturally flowing wells that although able to produce steadily the desired
flowrate, cannot start production without some help. Those wells need a kick-off
operation after a shutdown in order to produce a steady flowrate. In this case an artificial
lift method can be used whenever necessary to kick-off the well. In certain cases, the
bottom hole flowing pressure may be sufficient only to produce the well at flowrates
smaller than the recommended or desired flowrate. In some cases, the bottom hole
flowing pressure may not be capable to produce any flowrate at all and the well is called a
dead well. In those two cases artificial lift methods can be used to achieve the
recommended flowrate. Finally, there are conditions when the bottom hole flowing
pressure is able to produce the fluids to the surface but the production is unsteady. In
those cases artificial lift methods can be used to stabilize the well.

Artificial lift is therefore the area of petroleum engineering which is related to the use of
technologies to promote an increase in the production rate of flowing oil or gas wells, to
put wells back into production or to stabilize production by using an external horsepower
source that helps the bottom hole flowing pressure to overcome the pressure drops in the
system downstream of the perforations or to use methods that reduce the pressure drop
in the production system by improving the multiphase flow conditions in the well.

4.2 ARTIFICIAL LIFT MECHANISMS

1) Rod Pumps
Sucker Rod Pumps, also called donkey pumps or beam pumps, are the most common
artificial-lift system used in land-based operations (Figure 4.1). A motor drives a
reciprocating beam, connected to a polished rod passing into the tubing via a stuffing box.
The sucker rod continues down to the oil level and is connected to a plunger with a valve.
On each upward stroke, the plunger lifts a volume of oil up and through the wellhead
discharge. On the downward stroke it sinks (it should sink, not be pushed) with oil
flowing through the valve. The motor speed and torque is controlled for efficiency and
minimal wear with a Pump off Controller (PoC). Its use is limited to shallow reservoirs
down to a few hundred meters, and flows up to about 40 liters (10 gal) per stroke.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 67


Advanced Production Engineering 2019/2020 Academic Year

Figure 4.1: Sucker rod pump

Sucker Rod Pump is the most commonly used method worldwide. It is used massively in
US and in significant number of wells in Russia, China and South America. It is simple to
operate and inexpensive compared to other methods, but usually has lower pumping
capacity. It has problems with high gas liquid ratio (GLR), small tubing, deep wells and
crooked or deviated wells. It can achieve very low bottom hole flowing pressure making it
ideal for very depleted reservoirs.

Sucker Rod Pump


– The oldest and most commonly used form of artificial lift;
– Familiar to most engineers and operators;
– Allow very low fluid levels (low bottom hole flowing pressure);
– Low capital investment for low production at shallow to medium depths;
– High investment for high flow rates in deep wells;
– Adaptable to scale and corrosion problems;
– Limitation with casing size;
– Not suitable for crooked holes
– May have problems with sand production, high viscosity fluids and high GLR wells
– Easily installed in remote locations with an internal combustion engine.

2) Downhole Pumps
Downhole pump insert the whole pumping mechanism into the well. In modern
installations, an Electrical Submersible Pump (ESP) is inserted into the well. Here the
whole assembly consisting of a long narrow motor and a multiphase pump, such as a PCP
(progressive cavity pump) or centrifugal pump, hangs by an electrical cable with tension
members down the tubing (See Figure 4.2). ESPs works in deep reservoirs, but lifetime is
sensitive to contaminants such as sand, and efficiency is sensitive to GOR (Gas Oil Ratio)
where gas over 10% dramatically lowers efficiency.
Electrical Submersible Pump
– Second most commonly used method worldwide (+100,000 wells);
– Used massively in Russia and in significant number of wells in US;
– Responsible for the highest amount of total fluids produced (oil and water) by any
artificial lift method and an ideal method for high water cut wells;
– Problems with sand production, high GLR and high bottom hole temperatures.
– very high flow rates can be achieved;

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 68


Advanced Production Engineering 2019/2020 Academic Year

– do not require much space at the platform deck;


– very efficient to pump single phase low viscosity fluids;
– not very good at pumping very viscous fluids, high gas liquid ratio mixtures;
– very complex systems and the reliability is significantly smaller than gas lift. When
something serious goes wrong with the system, the well stop producing till the next
workover job.

Figure 4.2: Electrical submersible pump system

Progressing Cavity Pump (Figure 4.3)


– Less expensive than other methods and tolerant to solids and sand production;
– Usually has a stator made of an elastomer. Stator material is sensitive to oil
composition (aromatics) and temperature.
– May have problems with deviated and crooked wells.
– Applicable for low to medium flowrates.

Hydraulic Jet Pumping System


This method use a high pressure power fluid to drive a downhole turbine pump or flow
through a venturi or jet, creating a low pressure area which produces an increased
drawdown and inflow from the reservoir (Figure 4.4). It is very beneficial for high-
viscosity crudes where blending with a light power fluid can help in the processing of
heavy crudes. Hydraulic Jet Pumping system:
– Is very simple method;
– Has the biggest depth application range;
– Uses a light oil as the power fluid making it very interesting method for high
viscosity fluids;
– Requires an injection of 5 to 7 times the volume of liquid produced;
– Has no moving parts, applicable to deviated and crooked wells;
– Pump can be back-circulated to the surface.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 69


Advanced Production Engineering 2019/2020 Academic Year

Figure 4.3: Progressive cavity pump system

Figure 4.4: Hydraulic jet pumping system.

3) Gas Lift
Gas Lift injects gas into the well flow. The downhole reservoir pressure falls off to the
wellhead due to the counter pressure from weight of the oil column in the tubing. For
example, a 150 MPa reservoir pressure at 1600 meters will fall to zero wellhead pressure

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 70


Advanced Production Engineering 2019/2020 Academic Year

if the specific gravity is 800 kg/m2(0.8 times water). By injecting gas into this oil, the
specific gravity is lowered and the well will start to flow. Typically gas in injected between
casing and tubing, and a release valve on a gas lift mandrel is inserted in the tubing above
the packer (Figure 4.5). The valve will open at a set pressure to inject lift gas into the
tubing. Several mandrels with valves set at different pressure ranges can be used to
improve lifting and start up.

Figure 4.5: Gas lift system.


Continuous Gas Lift
– Very flexible and service through wireline is inexpensive. Tolerant to sand and solids
production, tolerant to high GLR, tolerant with deviated wells;
– Requires a stable source of high pressure gas and loses efficiency as bottom hole
flowing pressure declines and may require the use of another artificial lift method as
reservoir ages;
– The high reliability of the method makes it the first choice for offshore production
instead of ESP.

4) Plunger Lift
Plunger lift system (Figure 4.6) is normally used on low pressure gas wells with some
condensate, oil or water, or high gas ratio oil wells. In this case the well flow conditions
can be such that liquid starts to collect downhole and eventually blocks gas so that the
well production stops. In this case, a plunger with an open/close valve can be inserted in
the tubing. A plunger catcher at the top opens the valve and can hold the plunger, while
another mechanism downhole will close the valve. The cycle starts with the plunger
falling into the well with its valve open. Gas, condensate and oil can pass though the
plunger until it reaches bottom. There the valve is closed, now with a volume of oil,
condensate or water on top. Gas pressure starts to accumulate under the plunger and
after some time pushes the plunger upwards, with liquid on top, which eventually flows
out of the wellhead discharge. When the plunger reaches the wellhead plunger catcher,

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 71


Advanced Production Engineering 2019/2020 Academic Year

the valve opens and allows gas to flow freely for some time while new liquid collects at the
bottom. After some preset time the catcher will release the plunger, and the cycle repeats.

Figure 4.6: Plunger lift system.


Plunger Lift
– Can be used in combination with intermittent gas lift to reduce liquid fall back or by
itself in wells that present casing heading as a form of enhanced natural flow. Very
useful in dewatering applications

Some more advantages and disadvantages of the major artificial lift methods are given in
Tables 4.1 and 4.2.
Tables 4.1: Advantages of major artificial lift methods.

Tables 4.2: Disadvantages of major artificial lift methods.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 72


Advanced Production Engineering 2019/2020 Academic Year

4.3 SELECTION OF ARTIFICIAL LIFT CRITERIA


Selection of artificial lift system is a decision that takes into consideration several factors
such as: reservoir characteristics, fluid properties, well geometry and completion, field
location, operational problems, available equipment providers and type of energy
available (natural gas, mains electricity, diesel, etc).

Some of the well and reservoir characteristics include:


(i) Production casing size.
(ii) Maximum size of production tubing and required (gross) production rates.
(iii) Annular and tubing safety systems.
(iv) Producing formation depth and deviation (including doglegs, both planned and
unplanned).
(v) Nature of the produced fluids (gas fraction and sand/wax/asphaltene production).
(vi) Well inflow characteristics. A “straight line” inflow performance relationship
associated with a dead oil is more favourable than the curved “Vogel” relationship
found when well inflow takes place below the fluid’s bubble point. Figure 4.7
shows that reducing the flowing bottom hole pressure from 2500 to 500 psi
increases the well production rate by 125% for the dead oil. This is more than
double the 60% increase expected for the same reduction in bottom hole pressure
if a “Vogel” type inflow relationship is followed with a well producing below the
bubble point.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 73


Advanced Production Engineering 2019/2020 Academic Year

Figure 4.7: Influence of fluid in flow performance on production increase achieved when
well drawdown is increased

Field location
For offshore production, several other constraints must be considered when selecting an
artificial lift method:
– Deck space is limited and expensive
– Weight of artificial lift equipment that can be installed
– Energy available
– Distance from the wellhead to surface facilities (will determine minimum wellhead
flowing pressure (required for a give production rate)
– Workover operations and costs
– Type of well completion (subsea or dry completion)

For subsea wells, artificial lift methods that transmits energy mechanically (beam pump,
PCP, etc...) are not viable. Methods that use hydraulic, pneumatic or electric energy (gas
lift, hydraulic pump, electrical submersible pump) can be applied. Adaptation of gas lift
and hydraulic pumping to offshore subsea completion is straight forward and gas lift is
routinely used as the main and default artificial lift method for offshore subsea wells.
Adaptation of ESP for subsea wells was not direct and involved the development of
technology that allows the setup of underwater electric power distribution system and
wet electric connections. Those steps have been successfully achieved and today subsea
ESP is a proven technology.

For deep-water and ultra-deep- water applications, due to the high flow rates expected to
make production viable, only gas lift and ESP are considered as traditional artificial lift
methods. For this scenario both methods have advantages and disadvantages:

Operational problems
(i) Some forms of artificial lift e.g. gas lift are intrinsically more tolerant to solids
production (sand and/or formation fines) than other forms e.g. centrifugal pumps.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 74


Advanced Production Engineering 2019/2020 Academic Year

(ii) The formation of massive organic and inorganic deposits - paraffins, asphaltenes,
inorganic scales and hydrates. These are often preventable by treatment with suitable
inhibitors.
(iii) Producing velocities (erosion/corrosion).

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 75


Advanced Production Engineering 2019/2020 Academic Year

CHAPTER 5

SEPARATION
5.1 INTRODUCTION
Nature spends millions of years separating the oil/water in the reservoir. Production
engineers spend minutes mixing it up in the well bore which forces facility engineers to
spend countless hours and dollars designing and building equipment to separate the
phases all over again. Well produced fluids are a complex mixture of different
hydrocarbons components with different densities, vapour pressures and other physical
properties. Thus the well-stream may consist of crude oil, gas, condensates, water and
various contaminants. Depending on the mixture composition and the pressure and
temperature, the mixture can come in the following states: single phase liquid, single
phase gas and gas–liquid mixture. The purpose of the separators is to split the flow into
desirable fractions.

The physical separation of the gas-liquid mixture is one of the basic operations of the
processing of produced fluids. The oil-gas separator, mechanically separates the streams
of liquid and gas at a certain pressure and temperature. It is usually the first initial
processing vessel in a treatment facility. Separators must be properly designed. Poor
performance can reduce the capacity of the entire facility since downstream vessels and
equipment will be affected by the quality of the separated streams coming from the
separator. All separators are sized in accordance to the same principles and procedures.
Two Phase Separators separate gas from the total liquid (usually used for the first
separation stages) whereas three phase separators separate gas, oil and water (usually
used in the last separation stage).

5.2 TEST SEPARATORS AND WELL TEST


Test separators are used to separate the well flow from one or more wells for analysis and
detailed flow measurement. In this way, the behaviour of each well under different
pressure flow conditions can be determined. This normally takes place when the well is
taken into production and later at regular intervals, typically 1-2 months, measure the
total and component flow rates under different production conditions. Also undesirable
behaviour such as slugging or sand can be determined. The separated components are
also analyzed in the laboratory to determine hydrocarbon composition of the gas oil and
condensate. The test separator can also be used to produce fuel gas for power generation
when the main process is not running. In place of a test separator one could also use a
three phase flow meter to save weight.

5.3 PRODUCTION SEPARATORS


The main separators are gravity type. Figure 5.1 presents the main components around
the first stage separator. The production choke reduces pressure to the high pressure
(HP) manifold and first stage separator to about 3-5 MPa (30-50 times atmospheric
pressure). Inlet temperature is often in the range of 100-150 degrees C. The well stream is
colder due to subsea wells and risers. The pressure is often reduced in several stages;
here three stages are used, to allow controlled separation of volatile components. The
purpose is to achieve maximum liquid recovery and stabilized oil and gas, and separate
water. A large pressure reduction in a single separator will cause flash vaporization
leading to instabilities and safety hazards.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 76


Advanced Production Engineering 2019/2020 Academic Year

(a) (b
Figure 5.1: First stage separator (a) flow and process lines; (b) main
) components.

In the first stage separator, the water content is typically reduced to less than 5%.
Retention period is typically 5 minutes, allowing the gas to bubble out, water to settle at
the bottom and oil to be taken out in the middle. In this platform (Figure 5.1b) the water
cut (percentage water in the well flow) is almost 40% which is quite high. At the crude
entrance there is a baffle slug catcher that will reduce the effect of slugs (large gas bubbles
or liquid plugs). However some turbulence is desirable as this will release gas bubbles
faster than a laminar flow.

At the end there are barriers up to a certain level to keep back the separated oil and water.
The main control loops are the oil level control loop (EV0101 20 in Figure 5.1a)
controlling the oil flow out of the separator, and the gas pressure loop at the top (FV0105
20 in Figure 5.1a). These loops are operated by the Control System. An important function
is also to prevent gas blow-by which happens when low level causes gas to exit via the oil
output causing high pressure downstream. There are generally many more instruments
and control devices mounted on the separator.

The liquid outlets from the separator are equipped with vortex breakers (Figure 5.2) to
reduce disturbance on the liquid table inside. This is basically a flange trap to break any
vortex formation and ensure that only separated liquid is tapped off and not mixed with
oil or water drawn in through the vortices.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 77


Advanced Production Engineering 2019/2020 Academic Year

Figure 5.2: Different types of vortex breakers.

Similarly the gas outlets are equipped with demisters (mist extractors), essentially filters
that will remove liquid droplets in the gas. The demisters maybe:

a) Wire mesh: Liquid droplets impinge the wires and coalesce (Figure 5.3). Work on a
certain gas velocity range. At too low velocity, the droplets do not coalesce while at too
high velocity the gas can re-entrain the coalesced droplets. It is used only when
plugging by solids is unlikely.

Figure 5.3: Wire mesh demisters.

b) Vanes: Force the gas to flow in laminar regime between parallel channels with
directional changes. Liquid droplets impinge the walls of the channel, coalesce and
drop into the liquid collection area (Figure 5.4).

Figure 5.4: Vane demisters.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 78


Advanced Production Engineering 2019/2020 Academic Year

c) Centrifugal: Centrifugal forces cause the liquid to be separated. It is very efficient and
less susceptible to plugging. It is very sensitive to flow rate variation. There is large
pressure drop through the device. An example is presented in Figure 5.5.

Figure 5.5: Centrifugal demister.

d) Packing: Random packing can also be used as a coalescer for mist extraction (Figure
5.6).

Figure 5.6: Packing demister.

5.3.1 SECOND STAGE SEPARATOR


The second stage separator is quite similar to the first stage HP separator. It receives the
output of the first stage separator together with production from wells connected to the
Low Pressure manifold. The pressure will be around 1 MPa (10 atmospheres) and
temperature below 100 degrees C. The water content will be below 2%. An oil heater
could be located between the first and second stage separator to reheat the oil/water/gas
mixture. This will make it easier to separate out water when initial water cut is high and
temperature is low. The heat exchanger is normally a tube/shell type where oil passes
though tubes in a cooling medium placed inside an outer shell.

5.3.2 THIRD STAGE SEPARATOR


The final separator here is a two phase separator, also called a flash-drum (Figure 5.7).
The pressure is now reduced to about atmospheric pressure (100 kPa) so that the last
heavy gas components will boil out. In some processes where the initial temperature is

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 79


Advanced Production Engineering 2019/2020 Academic Year

low, it might be necessary to heat the liquid (in a heat exchanger) again before the flash
drum to achieve good separation of the heavy components. There are level and pressure
control loops. As an alternative, when the production is mainly gas, and remaining liquid
droplets have to be separated out, the two phase separator can be a Knock-Out Drum.

Figure 5.7: Third Stage Separator.

5.3.3 COALESCER
After the third stage separator, the oil can go to a coalescer (Figure 5.8) for final removal
of water. In this unit the water content can be reduced to below 0.1%. The coalescer is
completely filled with liquid: water at the bottom and oil on top. Inside electrodes form an
electric field to break surface bonds between conductive water and isolating oil in an oil-
water emulsion. The coalescer field plates are generally steel, sometimes covered with
dielectric material to prevent short circuits. The critical field strength in oil is in the range
0.2 to 2 kV/cm. Field intensity and frequency as well as the coalescer grid layout is
different for different manufacturers and oil types.

Figure 5.8: Coalescer.

5.3.4 ELECTROSTATIC DESALTER


If the separated oil contains unacceptable amounts of salts, it can be removed in an
electrostatic desalter. The salts, which may be Sodium, Calcium or Magnesium chlorides
comes from the reservoir water and is also dissolved in the oil. The desalters will be
placed after the first or second stage separator depending on Gas Oil Ratio (GOR) and
Water cut.

5.3.5 WATER TREATMENT


When the water cut is high, there will be a huge amount of produced water. For example, a
water cut of 40% can give a water production of about 4000 cubic meters per day (4
million liters) that must be cleaned before discharge to sea. Often this water contains sand

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 80


Advanced Production Engineering 2019/2020 Academic Year

particles bound to the oil/water emulsion. The environmental regulations in most


countries are quite strict, as an example, in the North-East Atlantic an oil in water
discharged to sea of 40 mg/liter (ppm) is allowed. It also places limits on other forms of
contaminants. Figure 5.9 shows a typical water treatment system. Water from the
separators and coalescers first goes to a sand cyclone, which removes most of the sand.
The sand is further washed before it is discharged.

Figure 5.9: Water treatment system.

The water then goes to a hydrocyclone, a centrifugal separator that will remove oil drops.
The hydrocyclone creates a standing vortex where oil collects in the middle and water is
forced to the side. Finally the water is collected in the water de-gassing drum. Dispersed
gas will slowly rise to the surface and pull remaining oil droplets to the surface by
flotation. The surface oil film is drained, and the produced water can be discharged to sea.
Recovered oil in the water treatment system is typically recycled to the third stage
separator.

5.4 GAS TREATMENT AND COMPRESSION


The gas train consists of several stages, each taking gas from a suitable pressure level in
the production separator’s gas outlet, and gas from other stages (See Figure 5.10).
Incoming gas (on the right) is first cooled in a heat exchanger. It then passes through the
scrubber to remove liquids and goes into the compressor.

Figure 5.10: Gas treatment and compression system.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 81


Advanced Production Engineering 2019/2020 Academic Year

The anti surge loop (thin orange line) and the surge valve (UV0121 23) allows the gas to
recirculate. The components are described below.

5.4.1 HEAT EXCHANGERS


For the compressor operate in an efficient way, the temperature of the gas should be low.
The lower the temperature is the less energy will be used to compress the gas for a given
final pressure and temperature. However both gas from separators and compressed gas
are relatively hot. Temperature exchangers of various forms are used to cool the gas. Plate
heat exchangers (Figure 5.11a) consist of a number of plates where the gas and cooling
medium pass between alternating plates in opposing directions. Tube and shell
exchangers (Figure 5.11b) place tubes inside a shell filled with of cooling fluid. The cooling
fluid is often pure water with corrosion inhibitors.

Figure 5.11: Heat exchangers (a) Plate type (b) Tube and shell type.

When designing the process it is important to plan the thermal energy balance. Heat
should be conserved e.g. by using the cooling fluid from the gas train (which is hot) to
reheat oil in the oil train. Excess heat is disposed e.g. by sea water cooling. However hot
seawater is extremely corrosive, so materials with high resistance to corrosion, such as
titanium must be used.

5.4.2 SCRUBBERS AND REBOILERS


The separated gas may contain mist and other liquid droplets. Liquid drops of water and
hydrocarbons also form when the gas is cooled in the heat exchanger, and must be
removed before it reaches the compressor. If liquid droplets enter the compressor they
will erode the fast rotating blades. A scrubber is designed to remove small fractions of
liquid from the gas. Various gas drying equipments are available, but the most common
suction (compressor) scrubber is based on dehydration by absorption in Tri Ethylene
Glycol (TEG). The scrubber consists of many levels of glycol layers (Figure 5.12). A large
number of gas traps (enlarged detail) force the gas to bubble through each glycol layer as
it flows from the bottom up each division to the top.

Lean glycol is pumped in at the top, from the holding tank. It flows from level to level
against the gas flow as it spills over the edge of each trap. During this process it absorbs

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 82


Advanced Production Engineering 2019/2020 Academic Year

liquids from the gas and comes out as rich glycol at the bottom. The holding tank also
functions as a heat exchanger for liquid from and to the reboilers.

Figure 5.12: Scrubber and reboiler system.

The glycol is recycled by removing the absorbed liquid. This is done in the reboiler, which
is filled with rich glycol and heated to boil out the liquids at temperature of about 130-180
°C (260-350°F) for a number of hours. Usually there is a distillation column on the gas
vent to further improve separation of glycol and other hydrocarbons. For higher capacity
there are often two reboilers which alternate between heating rich glycol and draining
recycled lean glycol.

5.4.3 COMPRESSOR ANTI SURGE AND PERFORMANCE


Several types of compressors are used for gas compression, each with different
characteristics such as operating power, speed, pressure and volume:
• Reciprocating Compressor that use a piston and cylinder design with 2-2 cylinders
are built up to about 30 MW power, around 500-1800 rpm (lower for higher
power) with pressure up to 5MPa (500 bars). Used for lower capacity gas
compression and high reservoir pressure gas injection (see Figure 5.13).
• Screw compressors are manufactured up to several MW of power, synchronous
speed (3000/3600 rpm) and pressure up to about 2.5 MPa (25 bar). Two counter
rotating screws with matching profiles provide positive displacement and a wide
operating range. (see Figure 5.14). Typical use is natural gas gathering.

Figure 5.13: Reciprocating Compressor Figure 5.14: Screw Compressor

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 83


Advanced Production Engineering 2019/2020 Academic Year

• Axial blade and fin type compressors with up to 15 wheels provide high volumes at
relatively low pressure differential (discharge pressure 3-5 times inlet pressure),
speeds of 5000-8000 rpm, and inlet flows to 200.000 m3/hour. Applications
include air compressors and cooling compression in LNG plants.
• The larger oil and gas installations use Centrifugal compressors with 3-10 radial
wheels, 6000 – 20000 rpm (highest for small size), up to 80 MW load at discharge
pressure of up to 50 bars and inlet volumes of up to 500.000 m3/hour. Pressure
differential up to 10.

Figure 5.15: Axial blade and fin compressor Figure 5.16: Centrifugal Compressor

Most compressors will not cover the full pressure range efficiently. Therefore
compression is divided into several stages to improve maintenance and availability.
Compressors are driven by gas turbines or electrical motors (for lower power also
reciprocating engines, steam turbines are sometimes used if thermal energy is available).
Often several stages in the same train are driven by the same motor or turbine.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 84


Advanced Production Engineering 2019/2020 Academic Year

CHAPTER 6

PRESSURE DROP IN PIPING

6.1 INTRODUCTION

Piping design in production facilities involves the selection of a pipe diameter and a wall
thickness that is capable of transporting fluid from one piece of process equipment to
another, within the allowable pressure drop and pressure rating restraints imposed by
the process. The first step in being able to make these changes is to understand how
pressure drops in these lines are calculated. While this chapter emphasizes piping that
exists within a facility, the concepts included on pressure drop are equally valid for
determining the pressure drop in flowlines, pipelines, gas transmission lines, etc.

This chapter first introduces the basic principles for determining pressure drops in piping
and then discusses the flow equations for liquid flow, compressible flow, and two-phase
flow. Finally, it shows how to calculate pressure drop in valves and fittings when using the
various flow equations. The last part of this chapter includes some example calculations
for determining the pressure drop in various types of pipe.

6.2 BASIC PRINCIPLES

Reynolds Number

The Reynolds number is a dimensionless parameter that relates the ratio of inertial forces
to viscous forces. It can be expressed by the following general equation:

DV
Re = (6.1)

where Re = Reynolds number


ρ = density, lb/ft3
D = pipe ID, ft.
V= flow velocity, ft/sec
µ = viscosity, lb/ft-sec

The Reynolds number can be expressed in more convenient terms. For liquids, the
equation becomes:

DV
Re = 7,738 (6.2)

Ql
Re = 92.1 (6.3)
D

where Re = Reynolds number


µ = viscosity, cp
D= pipe ID, in.
Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 85
Advanced Production Engineering 2019/2020 Academic Year

V= velocity, ft/sec
γ = specific gravity of liquid relative to water
Ql = liquid flow rate, bpd

For gas, the equation becomes:

Qg
Re = 20,100 (6.4)
D

where Re = Reynolds number


µ = viscosity, cp
D= pipe ID, in.
γ = specific gravity of gas at standard conditions (air = 1)
Qg = liquid flow rate, MMscfd

Flow Regimes
Flow regimes describe the nature of fluid flow. There are two basic flow regimes for flow
of a single-phase fluid: laminar flow and turbulent flow. Laminar flow is characterized by
little mixing of the flowing fluid and a parabolic velocity profile. Turbulent flow involves
complete mixing of the fluid and a more uniform velocity profile. Laminar flow has been
shown by experiment to exist at Re < 2,000 and turbulent flow at Re > 4,000. Reynolds
numbers between 2,000 and 4,000 are in a transition zone, and thus the flow may be
either laminar or turbulent.

Bernoulli's Theorem

It is customary to express the energy contained in a fluid in terms of the potential energy
contained in an equivalent height or "head" of a column of the fluid. Using this convention,
Bernoulli's theorem breaks down the total energy at a point in terms of:

i) The head due to its elevation above an arbitrary datum of zero potential energy.
ii) A pressure head due to the potential energy contained in the pressure in the fluid
at that point.
iii) A velocity head due to the kinetic energy contained within the fluid.

Assuming that no energy is added to the fluid by a pump or compressor, and that the fluid
is not performing work as in a steam turbine, the law of conservation of energy requires
that the energy at point "2" in the piping system downstream of point "1" must equal the
energy at point "1" minus the energy loss to friction and change in elevation. Thus,
Bernoulli's theorem may be written:

(Elevation Head)1 (Elevation Head)2


+ (Pressure Head)1 = + (Pressure Head) 2
+ (Velocity Head)1 + (Velocity Head) 2
+ Friction Head Loss)

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 86


Advanced Production Engineering 2019/2020 Academic Year

or

144P1 V12 144P2 V22


Z1 + + = Z2 + + + HL (6.5)
1 2g 2 2g

where Z = elevation head, ft


P = pressure, psi
ρ = density, lb/ft3
V = velocity, ft/sec
g = gravitation constant
HL= friction head loss, ft

Darcy's Equation
This equation, which is also sometimes called the Weisbach equation or the Darcy-
Weisbach equation, states that the friction head loss between two points in a completely
filled, circular cross section pipe is proportional to the velocity head and the length of pipe
and inversely proportional to the pipe diameter. This can be written:

fLV 2
HL = (6.6)
D2 g

where L = length of pipe, ft


D = diameter of pipe, ft
f = friction factor

Equations 6.5 and 6.6 can be used to calculate the pressure at any point in a piping system
if the pressure, flow velocity, pipe diameter, and elevation are known at any other point.
Conversely, if the pressure, pipe diameter, and elevations are known at two points, the
flow velocity can be calculated.

In most production facility piping systems the head differences due to elevation and
velocity changes between two points can be neglected. In this case Equation 6.5 can be
reduced to:


P1 − P2 = P = HL (6.7)
144

where ΔP = loss in pressure between points 1 and 2, psi

Substituting Equation 6.6 into Equation 6.7, we have:

fLV 2
P = (6.8)
144D 2 g

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 87


Advanced Production Engineering 2019/2020 Academic Year

Note: g = 32.2 ft/s2

D is in feet and so if the pipe diameter is in inches, divide by 12.

fLV 2 (12) fLV 2 (12)


P = =
144D2 g (144)(2)(32.2)D

fLV 2
P = 0.0013 (D in inches) (6.8*)
D

Moody Friction Factor


The Moody friction factor is determined from the Moody resistance diagram shown in
Figure 6.1. The friction factor is sometimes expressed in terms of the Fanning friction
factor, which is one fourth of the Moody friction factor. In some references the Moody
friction factor is used, in others, the Fanning friction factor is used. Care must be exercised
to avoid in advertent use of the wrong friction factor.

In general, the friction factor is a function of the Reynolds number, Re, and the relative
roughness of the pipe, ɛ/D. For Laminar flow, f is a function of only the Re:

64
f = (6.9)
Re

For turbulent flow, f is a function of both pipe roughness and the Reynolds number. At
high values of Re, f is a function only of ɛ/D.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 88


Advanced Production Engineering 2019/2020 Academic Year

Figure 6.1. Friction factor chart

Table 6.1 shows the relative roughness for various types of new, clean pipe. These values
should be increased by a factor of 2 – 4 to allow for age and use.

Table 6.1: Pipe roughness

6.3 FLUID FLOW EQUATIONS


Liquid Flow
It has already been shown that, pressure drop of liquid in a pipe can be found from:

fLV 2
P = 0.0013
D

where ρ = density, lb/ft3


L = length of pipe, ft
V = velocity, ft/sec
D = pipe diameter, in

The pressure drop can also be calculated using:

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 89


Advanced Production Engineering 2019/2020 Academic Year

fLQl2 
P = (11.5  10−6 ) (6.10)
D5

where ΔP = pressure drop, psi


f = Moody friction factor, dimensionless
L= length of pipe, ft
Ql = liquid flow rate, bpd
γ = specific gravity of liquid relative to water
D = pipe ID, in.

Derivation of Equation 6.10


fLV 2
P = 0.0013
D

ρ is in density, lb/ft3 ; V is in velocity, ft/sec; D in inches

Ql
V=
A
Ql is in ft3/sec and A in ft2

In case liquid flow rate Ql is known in bpd

ft 3 day hr
Q = Q1  5.61   = 6.49 10−5 Q1
barrel 24hrs 3,600s

In case A is known in inches


D 2
A=
(4)(144)

Ql (6.49 10−5 )(4)(144)Ql 0.0119Ql


V= = =
A D 2 D2
 = 62.4

fLV 2
Substituting the V and ρ terms in P = 0.0013 gives:
D

fLQl2 
P = (11.5  10−6 )
D5

The most common use of Equation 6.10 is to determine a pipe diameter for a given flow
rate and allowable pressure drop. However, first, a calculation of Reynolds number
(Equation 6.2 or 6.3) to determine the friction factor must be made. Since a Reynolds
number depends on the pipe diameter the equation cannot be solved directly. One method

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 90


Advanced Production Engineering 2019/2020 Academic Year

to overcome this disadvantage is to assume a typical friction factor of 0.025, solve


Equation 6.10 for diameter, compute a Reynolds number, and then compare the assumed
friction factor to one read from Figure 6.1. If the two are not sufficiently close, it is
possible to iterate the solution until convergence.

Figure 6.2 is a curve that can be used to approximate pressure drop or required pipe
diameter.

Figure 6.6: Pressure drop in liquid lines

It is based on an assumed friction factor relationship, which can be adjusted to some


extent for liquid viscosity. In an effort to void an iterative calculation, several empirical
formulas have been developed. The most common of these is the Hazen-Williams
formula, which can be expressed as follows:

1.85 1.85
 100   Qo 
H L = 0.00208   4.87  L (6.11)
 C  D 

H L = 0.015
(Q )(L )
1.85
(6.12)
(D )(C )
1
4.87 1.85

where HL = head loss due to friction, ft


L = length, ft
C = friction factor constant, dimensionless
=140 for new steel pipe
Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 91
Advanced Production Engineering 2019/2020 Academic Year

=130 for new cast iron pipe


=100 for riveted pipe
D = pipe ID, in.
Qo = liquid flow rate, gpm
Q1 = liquid flow rate, bpd

This equation is based on water flowing under turbulent conditions with a viscosity of
1.13 centipoise, which is the case for water at 60°F. Since water viscosity varies
appreciably from 32°F to 212°F, the friction factor can decrease or increase as much as
40% between the two temperature extremes.

The Hazen-Williams equation is frequently used for calculating pressure losses and line
capacities in water service. The discharge coefficient "C" must be carefully chosen to
reflect both fluid viscosity and pipe roughness in a used condition. A "C" factor of 90 to
100 in steel pipe is common for most produced liquid problems. Hazen-Williams factor
"C" must not be confused with the Moody friction factor "f," as these two factors are not
directly related to each other. Typical "C" factors for various types of pipe are shown in
Table 6.2.

Table 6.2: Hazen-Williams “C” Factors

Gas Flow
The Darcy equation assumes constant density over the pipe section between the inlet and
outlet points. While this assumption is valid for liquids, it is incorrect for pipelines flowing
gases, where density is a strong function of pressure and temperature. As the gas flows
through the pipe it expands due to the drop in pressure and thus tends to decrease in
density.
At the same time, if heat is not added to the system, the gas will cool, causing the gas to
tend to increase in density. In control valves, where the change in pressure is near
instantaneous, and thus no heat is added to the system, the expansion can be considered
adiabatic. In pipe flow, however, the pressure drop is gradual and there is sufficient pipe
surface area between the gas and the surrounding medium to add heat to the gas and thus
keep the gas at constant temperature. In such a case the gas can be considered to undergo
an isothermal expansion. On occasion, where the gas temperature is significantly different
from ambient, the assumption of isothermal (constant temperature) flow is not valid. In

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 92


Advanced Production Engineering 2019/2020 Academic Year

these instances greater accuracy can be obtained by breaking the line up into short
segments that correspond to only small temperature changes.

The general isothermal equation for the expansion of gas can be given by:
 
 
  (P1 ) − (P2 ) 
2 2
 144gA2 (6.13)
w =
2
 
  fL P1    P1
V 1  + 2 log e   
  D P2  

where w = rate of flow, lb/sec


g = acceleration due to gravity, ft/sec2
A = cross-sectional area of pipe, ft2
V 1 = specific volume of gas at upstream conditions, ft3/lb
f = friction factor
L = length, ft
D = diameter of pipe, ft
P1 = upstream pressure, psia
P2 = downstream pressure, psia

This equation assumes that:

i) No work is performed between points 1 and 2, i.e., there are no compressors or


expanders and no elevation changes.
ii) The gas is flowing under steady state conditions, i.e., no acceleration changes.
iii) The Moody friction factor, f, is constant as a function of length. There is some
change due to a change in Reynolds number, but this is quite small.

For practical pipeline purposes,


P1 fL
2 loge 
P2 D
With this assumption and substituting in Equation 6.13 for practical oil field units:

Qg2 ZT1 fL
P12 − P22 = 25.1 (6.14)
D5

where P1 = upstream pressure, psia


P2 = downstream pressure, psia
γ = gas specific gravity at standard conditions
Qg = gas flow rate, MMscfd
Z = compressibility factor for gas
T1 = flowing temperature, °R
f = Moody friction factor, dimensionless
D = pipe ID, in.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 93


Advanced Production Engineering 2019/2020 Academic Year

The "Z" factor will change slightly from point 1 to point 2. It is usually assumed to be
constant and is chosen for an "average" pressure of:

2 P1 P2 
 P1 + P2 −  (6.15)
3 P1 + P2 

Rearranging Equation 6.14 and solving for Qg we have:

( )
1
 D 5 P12 − P22  2
Q g = 0.199  (6.16)
 ZT1 fL 

An approximation of Equation 6.14 can be made when the change in pressure is less than
10% of the inlet pressure. If this is true we can make the assumption:

P12 − P22  2 P1 (P1 − P2 )

Substituting into Equation 6.14, we have:

 Qg2 ZT1 fL 
P = 12.6 5  (6.17)
 P1 D 

As was the case for liquid flow, in order to solve any of these equations for a pipe diameter
to handle a given flow and pressure drop, it is necessary to first guess the diameter and
then compute a Reynolds number to determine the friction factor. Once the friction factor
is known, a pipe diameter can be calculated and compared against the assumed number.
The process can be iterated until convergence.

Several empirical gas flow equations have been developed. These equations are patterned
after the general flow equation (Equation 6.16), but make certain assumptions so as to
avoid solving for the Moody friction factor. The three most common gas flow equations
are described in the following sections.

Figure 6.3 can be used to estimate the viscosity of a hydrocarbon gas at various conditions
of temperature and pressure if the specific gravity of the gas at standard conditions is
known. Again if the specific gravity of the gas at standard conditions is known the
compressibility factor can be found using the chart under Appendix A.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 94


Advanced Production Engineering 2019/2020 Academic Year

Figure 6.3: Hydrocarbon gas viscosity

Weymouth Equation
This equation is based on measurements of compressed air flowing in pipes ranging from
0.8 in. to 11.8 in. in the range of the Moody diagram where the ɛ/D curves are horizontal
(i.e., high Reynolds number). In this range the Moody friction factor is in dependent of the
Reynolds number and dependent upon the relative roughness. For a given absolute
roughness, ɛ, the friction factor is merely a function of diameter. For steel pipe the
Weymouth data indicate:

0.032
f = (6.18)
D1 3

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 95


Advanced Production Engineering 2019/2020 Academic Year

Substituting this into Equation 6.16, the Weymouth equation expressed in practical
oilfield units is:
1
 P 2 − P22  2
Q g = 1.11D 2.67  1  (6.19)
 ZT1 L 

where Qg = flow rate, MMscfd


D = pipe ID, in.
P1 and P2=pressure at points 1 and 2 respectively, psia
L = length of pipe, ft
γ = specific gravity of gas at standard conditions
T1 = temperature of gas at inlet, °R
Z = compressibility factor of gas

Assuming a temperature of 520°R (standard condition), a compressibility of 1.0 and a


specific gravity of 0.6 the Weymouth equation can also be written:

1
 P12 − P22  2
Qg = 865D 2.67
  (6.20)
 L 

where Qg = flow rate, scfd


D = pipe ID, in.
L = pipe length, miles

It is important to remember the assumptions used in deriving this equation and when
they are appropriate. Short lengths of pipe with high pressure drops are likely to be in
turbulent flow and thus the assumptions made by Weymouth are appropriate. Industry
experience indicates that Weymouth's equation is suitable for most piping within the
production facility. However, the friction factor used by Weymouth is generally too low
for large diameter or low velocity lines where the flow regime is more properly
characterized by the sloped portion of the Moody diagram.

Panhandle Equation
This equation was intended to reflect the flow of gas through smooth pipes and is a
reasonable approximation of partially turbulent flow behaviour. The friction factor can be
represented by a straight line of constant negative slope in the moderate Reynolds
number region of the Moody diagram.

A straight line on the Moody diagram would be expressed:

log f = n log Re + log C (6.21)

or

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 96


Advanced Production Engineering 2019/2020 Academic Year

C
f = (6.22)
Ren

The Panhandle A equation applies to Reynolds numbers in the 5 x 106 to 11 x 106 range
and assumes n = 0.146. The Panhandle B equation assumes more fully developed
turbulent flow (greater Reynolds number) and assumes a lower slope of n = 0.039.

Using this assumption and assuming a constant viscosity for the gas, the Panhandle A
equation can be written:
0.059
 P12 − P22 
Qg = 0.020E  0.853  D 2.62
 ZT1 L 

Panhandle B equation can be written:


0.51
 P12 − P22 
Qg = 0.028E  0.961  D 2.53
 ZT1 L 

where Qg = flow rate, MMscfd


D = pipe ID, in.
L = pipe length, miles
E = efficiency factor
= 1.0 for brand new pipe
= 0.95 for good operating conditions
= 0.92 for average operating conditions
= 0.85 for unfavourable operating conditions

In practice, the Panhandle equations are commonly used for large diameter, long pipelines
where the Reynolds number is on the straight line portion of the Moody diagram. It can be
seen that neither the Weymouth nor the Panhandle equations represent a "conservative"
assumption. If the Weymouth formula is assumed, and the flow is a moderate Reynolds
number, the friction factor will in reality be higher than assumed (the sloped line portion
is higher than the horizontal portion of the Moody curve), and the actual pressure drop
will be higher than calculated. If the Panhandle B formula is used and the flow is actually
in a high Reynolds number, the friction factor will in reality be higher than assumed (the
equation assumes the friction factor continues to decline with increased Reynolds number
beyond the horizontal portion of the curve), and the actual pressure drop will be higher
than calculated.

Spitzglass Equation
This equation is used for near-atmospheric pressure lines. It is derived directly from
Equation 6.16 by making the following assumptions:

 3.6  1 
i) f = 1 + + 0.03D  
 D  100 
ii) T = 520oR
Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 97
Advanced Production Engineering 2019/2020 Academic Year

iii) P1 =15 psi


iv) Z = 1.0
v) ΔP< 10% of P1

With these assumptions, and expressing pressure drop in terms of inches of water, the
Spitzglass equation can be written:
1
 2
 hw D 5 
Q g = 0.09  (6.25)
  3.6 
 L1 + 1 + D + 0.03D  
 

where Δhw = pressure loss, inches of water


D = pipe ID, in.
L = pipe length, ft

6.4 APPLICATION OF GAS FLOW EQUATIONS


The Weymouth and Spitzglass equations both assume that the friction factor is merely a
function of pipe diameter. Figure 6.4 compares the friction factors calculated from these
equations with the factor indicated by the horizontal line of the Moody diagram for two
different absolute roughnesses.

Figure 6.4: Friction factor vs. pipe diameter for three correlations.
In the small pipe diameter range (3-6 in.) all curves tend to yield identical results. For
large diameter pipe (10 in. and larger) the Spitzglass equation becomes overly
conservative. The curve is going in the wrong direction, thus the form of the equation
must be wrong. The Weymouth equation tends to become under-conservative with pipe
greater than 20 inches. Its slope is greater than the general flow equation with ɛ = 0.002
Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 98
Advanced Production Engineering 2019/2020 Academic Year

inch. This is merely a result of the way in which the Spitzglass and Weymouth equations
represent the Moody diagram.

The empirical gas flow equations use various coefficients and exponents to account for
efficiency and friction factor. These equations represent the flow condition upon which
they were derived, but may not be accurate under different conditions. Unfortunately,
these equations are often used as if they were universally applicable. The following
guidelines are recommended in the use of gas flow equations:

i) The general gas flow equation is recommended for most general usage. If it is
inconvenient to use the iterative procedure of the general equation and it is not
known whether the Weymouth or the Panhandle equations are applicable,
compute the results using both Weymouth and Panhandle equations and use the
higher calculated pressure drop.

ii) Use the Weymouth equation only for small-diameter, short-run pipe within the
production facility where the Reynolds number is expected to be high.

iii) Use the Panhandle equations only for large-diameter, long-run pipelines where the
Reynolds number is expected to be moderate.

iv) Use the Spitzglass equation for low pressure vent lines less than 12-inches in
diameter.

v) When using gas flow equations for old pipe, attempt to derive the proper efficiency
factor through field tests. Buildup of scale, corrosion, liquids, paraffin, etc. can have
a large effect on gas flow efficiency.

Two-Phase Flow
Two-phase flow of liquid and gas is a very complex physical process. Even when the best
existing correlations for pressure drop and liquid holdup are used, predictions may be in
error as much as ±20%. Nevertheless, as gas exploration and production have moved into
remote offshore, arctic, and desert areas, the number of two-phase pipelines has
increased.

To determine whether two-phase flow will exist in a pipeline, the expected flowing
pressure and temperature ranges in the line must be plotted on a phase diagram for the
fluid. Figure 6.5 shows that composition B will flow as a single-phase fluid as it enters the
pipeline. However, as the pressure drops it becomes a two-phase mixture through part of
the pipeline. On the other hand, composition A will flow as a single-phase (dense fluid or
gas) through the entire length of the line. Composition C will flow as a liquid throughout
the entire length of the line.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 99


Advanced Production Engineering 2019/2020 Academic Year

Figure 6.5: Typical hydrocarbon phase behaviour.

Horizontal Flow
When a gas-liquid mixture enters a pipeline, the two phases tend to separate with the
heavier liquid gravitating to the bottom. Figure 6.6 shows typical flow patterns in
horizontal two-phase pipe flow.

Figure 6.6:Two-phase flow patterns in horizontal flow.

The type of flow pattern depends primarily on the superficial velocities as well as the
system geometry and physical properties of the mixture. At very low gas-liquid ratios, the
gas tends to small bubbles that rise to the top of the pipe. As the gas-liquid ratio increases,
the bubbles become larger and eventually combine to form plugs. Further increases in the
gas-liquid ratio cause the plugs to become longer, until finally the gas and liquid phases
flow in separate layers; this is stratified flow. As the gas flow rate is increased, the gas-
liquid interface in stratified flow becomes wavy. These waves become higher with
increasing gas-liquid ratios, until the crest of the waves touches the top of the pipe to form
slugs of liquid which are pushed along by the gas behind them. These slugs can be several
hundred feet long in some cases. Further increases in the gas-liquid ratio may impart a

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 100
Advanced Production Engineering 2019/2020 Academic Year

centrifugal motion to the liquid and result in annular flow. At extremely high gas-liquid
ratios, the liquid is dispersed into the flowing gas stream.

Figure 6.7 shows how the flow regime for horizontal pipes depends primarily on the
superficial gas and liquid flow rates. Experience has shown that generally flow regime
maps such as Figure 6.7 are not very accurate, but they can be used as qualitative guides.

In most two-phase flow lines in the field, slug flow is predominant in level and uphill lines.
In downhill lines, stratified flow is predominant. However, if the slope of the downhill line
is not very steep and the gas velocity is high, slug flow may be observed. The criterion for
transition from stratified to slug flow in downhill lines is not well defined.

Figure 6.7: Flow regime for horizontal pipes.

Vertical Flow
The two-phase flow patterns in vertical flow are somewhat different from those occurring
in horizontal or slightly inclined flow. Vertical two phase flow geometries can be classified
as bubble, slug-annular, transition, and annular-mist, depending on the gas-liquid ratio.
All four flow regimes could conceivably exist in the same pipe. One example is a deep well
producing light oil from a reservoir that is near its bubble point. At the bottom of the hole,
with little free gas present, flow would be in the bubble regime. As the fluid moves up the
well, the other regimes would been countered because gas continually comes out of
solution as the pressure continually decreases. Normally flow is in the slug regime and
rarely in mist, except for condensate reservoirs or steam-stimulated wells. The different

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 101
Advanced Production Engineering 2019/2020 Academic Year

flow regimes are shown in Figure 6.8. Figure 6.9 gives approximate flow regimes as a
function of superficial gas and liquid flow rates.

Figure 6.8: Two-phase flow patterns in vertical low.

Figure 6.9: Vertical multiphase flow map.

These flow regimes are described below:

i) Bubble Flow: The gas-liquid ratio is small. The gas is present as small bubbles,
randomly distributed, whose diameters also vary randomly. The bubbles move at
different velocities depending upon their respective diameters. The liquid moves
up the pipe at a fairly uniform velocity, and except for its density, the gas phase has
little effect on the pressure gradient.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 102
Advanced Production Engineering 2019/2020 Academic Year

ii) Slug Flow: In this regime the gas phase is more pronounced. Although the liquid
phase is still continuous, the gas bubbles coalesce and form stable bubbles of
approximately the same size and shape, which are nearly the diameter of the pipe.
They are separated by slugs of liquid. The bubble velocity is greater than that of the
liquid and can be predicted in relation to the velocity of the liquid slug. There is a
film of liquid around the gas bubble. The liquid velocity is not constant; whereas
the liquid slug always moves upward (in the direction of bulk flow), the liquid in
the film may move upward, but possibly at a lower velocity, or it may even move
downward. These varying liquid velocities not only result in varying wall friction
losses, but also result in liquid holdup, which influences flowing density. At higher
flow velocities, liquid can even be entrained in the gas bubbles. Both the gas and
liquid phases have significant effects on pressure gradient.

iii) Transition Flow: The change from a continuous liquid phase to a continuous gas
phase occurs in this region. The liquid slug between the bubbles virtually
disappears, and a significant amount of liquid becomes entrained in the gas phase.
In this case, although the effects of the liquid are significant, the gas phase is
predominant. Transition flow is also known as "churn flow."

iv) Annular-Mist Flow: The gas phase is continuous. The bulk of the liquid is
entrained and carried in the gas phase. A film of liquid wets the pipe wall, but its
effects are secondary. The gas phase is the controlling factor.

Pressure Drop in Two-Phase Flow


Pressure drop in two-phase flow is the sum of the pressure drop due to acceleration,
friction losses, and elevation changes. In most pipelines, the pressure loss due to
acceleration is small. Pressure drop due to friction is typically several times larger in two-
phase flow than the sum of the pressure drops of the equivalent two single phases. The
additional frictional pressure drop in two-phase flow is attributed to irreversible energy
transfer between phases at the interface and to the reduced cross-sectional area available
for the flow to each phase.

Pressure drop due to elevation changes is also significant in two-phase flow. In an uphill
line, the pressure drop due to elevation change is merely the average density of the two
phase mixture in the uphill line multiplied by the change in elevation. Since the average
density depends on the liquid holdup, the static head disadvantage in an uphill line also
depends on the average liquid holdup for the segment.

In cross-country lines that consist of a number of uphill and downhill segments, the worst
case for pressure drop occurs at low gas flow rates, where each uphill segment fills with
liquid. That is, in these segments the liquid holdup approaches the volume of the pipe and
the gas is in bubble flow. When this happens, the pressure drop in each uphill segment is
given by:

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 103
Advanced Production Engineering 2019/2020 Academic Year

Pz  0.43Z n (6.26)

where ΔPZ = pressure drop due to elevation increase in the segment, psi
ΔZn = increase in elevation for segment n, ft
γ = specific gravity of the fluid in the segment relative to water

In downhill lines, flow is normally stratified, and the liquid flows as fast or faster than the
gas. The depth of the liquid layer adjusts to the critical depth at which the static head
advantage due to the downhill run balances the frictional losses and the pressure loss can
be considered to be zero. If slug flow exists in the downhill segment due to a slight slope
and/or high gas velocity, the static head advantage will generally be less than the
frictional pressure drop, resulting in a net pressure loss. In cross-country lines the static
head advantage is normally neglected, and in hand calculations no pressure recovery due
to the decrease in elevation is considered.

The net effect of a pressure loss due to an increase in elevation on each uphill segment
and no pressure recovery on each downhill segment can be quite severe. In a single-phase
line only the net change in elevation from the beginning of the line to the end of the line
need to be considered.
In a two-phase line Equation 6.26 states the pressure lost in each uphill segment, but this
is not balanced by a pressure gain in each downhill segment. Thus, the pressure lost due
to elevation changes is the sum of the pressure lost in each uphill segment. It is possible in
hilly terrain to have a line that flows from a high point to a lower elevation, but which,
because it crosses a valley, will still have a loss in pressure due to elevation changes.
Hand calculation methods have been developed to calculate pressure drops in two-phase
flow. One equation presented in the American Petroleum Institute's Recommended
Practice entitled "Design and Installation of Offshore Production Piping Systems" is:

3.4  10 −6 fLW 2
P = (6.27)
m D5

where L = length, ft
W = rate of flow of liquid and vapour, lb/hr
ρm = density of the mixture, lb/ft3
d = pipe ID, in.

For f = 0.015

5  10 −8 LW 2
P = (6.28)
m D5

This equation is derived from the general equation for isothermal flow by making the
following assumptions:
i) ΔP is less than 10% of inlet pressure

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 104
Advanced Production Engineering 2019/2020 Academic Year

ii) Bubble or mist flow exists


iii) No elevation changes

The rate of flow of the mixture to use in this equation can be calculated as follows:

W = 3,180Qg  g + 14.6Ql  l w (6.29)

where Qg = gas flow rate, MMscfd


Ql= liquid flow rate, bpd
γg = specific gravity of gas at standard conditions (air=1)
γlw = specific gravity of liquid relative to water

The density of the mixture to use in Equations 6.27 and 6.28 is given by:

12,409 l w P + 2.7 R g P
m = (6.30)
198.7 P + RTZ

where P = operating pressure, psia


R = gas/liquid ratio, ft3/bbl
T = operating temperature, °R
Z = gas compressibility factor

6.5 HEAD LOSS IN VALVES AND PIPE FITTINGS


In many piping problems, especially those associated with offshore production facilities
where space limitation are important, the pressure drop through valves, bends, tees,
enlargements, contractions, etc. becomes very important. The three most common ways of
handling these additional pressure drops are by use of resistance coefficients, flow
coefficients, and equivalent lengths.

Resistance Coefficients
Darcy's equation can be rewritten as:

V2
H L = Kr (6.31)
2g
where Kr = resistance coefficient
= fL/D

Although Kr depends on the Reynolds number and surface roughness as well as on the
geometry of the elbow or couplings, this dependence is usually neglected. Approximate
values of Kr are given in Table 6.3 for various pipe fittings. Figures 6.10 and 6.11 show
resistance coefficient for sudden contractions and expansions, and for entrances and exits.
Not that the friction factor (fT) in Figure 6.10 assumes fully turbulent flow through
standard tees and elbows.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 105
Advanced Production Engineering 2019/2020 Academic Year

The total head loss for the entire piping system can be determined from the following
equation:
V2
H L =  Kr (6.32)
2g

Table 6.3: Resistance Coefficient for Pipe Fittings

Figure 6.10: Representative resistance coefficients (K) for fittings.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 106
Advanced Production Engineering 2019/2020 Academic Year

Figure 6.11: Resistance in pipe due to sudden enlargements and contractions.

Flow Coefficient
The pressure drop characteristics of control valves are often expressed in terms of Cv, the
flow coefficient. The flow coefficient is measured experimentally for each valve or fitting
and is equal to the flow of water, in gpm, at 60°F for a pressure drop of one psi. It can be
shown from Darcy's equation that with Cv measured in this manner:

29.9d 2
Cv = (6.33)
( fL / D )1 2

where D = fitting equivalent ID, ft


d = fitting equivalent ID, in.
L = fitting equivalent length, ft

It follows from the definition of Kr that the relationship between Cv and Kr is:

29.9d 2
Cv = (6.34)
(K r )1 2

The pressure drop for any valve or fitting for which Cv is known is derived as follows:

Q 
P = 8.5  10 − 4  l  (6.35)
 Cv 

where Ql = Liquid flow rate, bpd

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 107
Advanced Production Engineering 2019/2020 Academic Year

γ = Liquid specific gravity relative to water

Equivalent Length
It is often simpler to treat valves and fittings in terms of their equivalent length of pipe.
The equivalent length of a valve or fitting is the length of an equivalent section of pipe of
the same diameter that gives the same head loss. Total head loss or pressure drop is
determined by adding all equivalent lengths to the pipe length. The equivalent length, Le,
can be determined from Kr or Cv as follows:

Kr D
Le = (6.36)
f
Krd
Le = (6.37)
12 f
74.5d 5
Le = (6.38)
fC v2

Table 6.4 summarizes the equivalent length for various commonly used valves and
fittings.

Table 6.4: Equivalent Length of Valves and Fittings in Feet

Figures 6.12 and 6.13 show equivalent lengths of various fabricated bends.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 108
Advanced Production Engineering 2019/2020 Academic Year

Figure 6.12: Equivalent length of 90° bends.

Figure 6.13: Equivalent length of mitre bends.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 109
Advanced Production Engineering 2019/2020 Academic Year

EXAMPLE PRESSURE DROP CALCULATIONS

Example 6.1: Pressure Drop in Liquid Line

Given the data below solve for pressure drop in a 2-inch and 4-inch I.D. line using the
general equation and Hazen-Williams:

Flow rates: Condensate = 800 bpd


Water = 230 bpd
Specific gravity: Condensate = 0.87
Water = 1.05
Viscosity = 3 cp
Length = 7,000 ft
Inlet pressure = 900 psi
Temperature = 80oF

Solution
1. General Equation
Specific gravity of liquid:
Q Q 
l = c c + w w
QT QT

l =
(800)(0.87) + (230)(1.05)= 0.91
1,030 1,030

Using Equation 6.3:

 l Ql (92.1)(0.91)(1,030)
Re = 92.1 =
D 3D

ɛ = 0.004 (assume old steel)

Calculate the pressure drop using Equation 6.10:

fLQl2 
P = (11.5  10−6 )
D5

P =
(11.5  10 )(7,000)(1,030) (0.81) f
−6 2
=
77,725 f
5
D D5

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 110
Advanced Production Engineering 2019/2020 Academic Year

2. Hazen-William

Assume C = 120

Using Equation 6.12:

H L = 0.015
(Q )(L) = 0.015 (1,030) (7,000)
1.85
l
1.85

(D )(C )
4.87 1.85
(120) D 1.85 4.87

5,604
HL =
D 4.87

Calculate the pressure drop using Equation 6.7:

l ( l )(62.4)
P = HL = HL
144 144

P =
(0.91)(62.4)(5,604)
144D 4.87

For D = 2 in. ΔP = 74.7 psi


For D = 4 in. ΔP = 2.6 psi

Example 6.2: Pressure Drop in Gas Line

Given the data below:

Flow rates: Gas = 23 MMscfd


Gravity: Gas = 0.85
Length = 7,000 ft
Inlet pressure = 900 psi
Temperature = 80°F

Solve for pressure drop in a 4-inch and 6-inch I.D. line using the:
i) General equation
ii) Assumption of ΔP < 10% P1
iii) Panhandle B Equation
iv) Weymouth Equation.
Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 111
Advanced Production Engineering 2019/2020 Academic Year

Solution
1. General Equation
Using specific gravity of gas:

Find the gas viscosity using the temperature and pressure:

Gas viscosity = 0.013 (from Chart in Figure 6.3)

Calculate Reynolds number using Equation 6.4

Qg (20,100)(0.85)(23) 30,227,000


Re = 20,100 = =
D 0.013D D

ɛ = 0.004 (assume old steel)

Z = 0.67 (from Chart in Appendix A)

Using Equation 6.14:

Qg2 ZT1 fL
P12 − P22 = 25.1
D5

P −P
2 2
= 25.2
(0.85)(23) (0.67)(540)(7,000) f
2

1 2
D5
2.87  1010 ( f )
P12 − P22 =
D5

2. Approximate Equation
Calculate the pressure drop using the approximate equation (Equation 6.17):

 Qg2 ZT1 fL 
P = 12.6 5 
 P1 D 

P =
(12.6)(0.85)(23)2 (0.67)(540)(7,000) f
900D 5

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 112
Advanced Production Engineering 2019/2020 Academic Year

1.59  107 ( f )
P =
D5

For D = 4 in. ΔP = 308 psi


For D = 6 in. ΔP = 37psi

3. Panhandle B Equation

7,000
L= = 1.33 miles
5,280

E = 0.95 (assumed)

Using Panhandle Equation B:

0.51
 P2 − P2 
Qg = 0.028E  01.961 2  D 2.53
 ZT1 L 

(900)2 − P22
0.51
 
23 = (0.028)(0.95)  D 2.53
 (0.85) (0.67)(540)(1.33)
0.961

1.96
810  103 P22  
4.96
23 1
− =
412 412  (0.028)(0.95)   D 

235  106
P = 810  10 −
2
2 3

D 4.96

4. Weymouth Equation
Using Weymouth Equation (Equation 6.19)

1
 P 2 − P22  2
Q g = 1.11D 2.67  1 
 ZT1 L 
(900)2 − P22
0.5
 
23 = 1.11D 2.67
 
 (0.85)(0.67)(540)(7,000)

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 113
Advanced Production Engineering 2019/2020 Academic Year

2
810  103  23   1 
5.33
P22
− =   
2.2  10 2.2  10  (1.11)   D 
6 6

 1 
810  103 − P22 = 9.44  108  5.33 
D 

9.44  108
P22 = 810  103 −
D5.33

Example 6-3: Pressure Drop in Two-Phase Line


Given: Same conditions as Examples 6.1 and 6.2, solve for the pressure drop with both
liquid and gas flow in a single 4-in., 6-in., or 8-in. line.

Solution
Specific gravity of liquid = 0.91 (from before)
Z = 0.67 (from Appendix A)

Calculate flow rate using Equation 6.29.

W = 3,180Qg  g + 14.6Ql  l w

W = 3,180(23)(0.85) + 14.6(1,030)(0.91)

W = 75,854 lb/hr

Calculate the gas liquid ratio:

23,000,000
R= = 22,330 ft3/bbl
1,030

Density of mixture (compute at 900 psi):

12,409 l w P + 2.7 R g P
m =
198.7 P + RTZ

m =
(12,409)(0.91)(900) + 2.7(22,33030)(0.85)(900)
(198.7 )(900) + (22,330)(540)(0.67)
 m = 6.93 lb/ft3

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 114
Advanced Production Engineering 2019/2020 Academic Year

Calculate the pressure drop using Equation 6.28:

5  10 −8 LW 2
P =
m D5

P =
(5 10 )(7,000)(74,854)
−8 2

(6.93)D 5

2.83  10 −5
P =
D5

For D = 4 in. ΔP = 276 psi


For D = 6 in. ΔP = 36psi
For D = 8 in. ΔP = 9 psi

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 115
Advanced Production Engineering 2019/2020 Academic Year

REFERENCES AND SUGGESTED READINGS

1. Ahmed, T., (2006), “Reservoir Engineering Handbook”, Gulf Professional Publishing


(an imprint of Elsevier), 3rd Edition, 1377 pp.
2. Anon, “Material on Production Technology II” Institute of Petroleum Engineering,
Heriot-Watt University.
3. Boyun, G, William, C. L. and Ali, G. (2007), “Petroleum Production Engineering – A
Computer-Assisted Approach” Gulf Professional Publishing, Houston, TX, pp. 3/30 –
3/43.
4. Havard, D. (2013), “Oil and Gas Production Handbook – An Introduction to Oil and Gas
Production, Transport, Refining and Petrochemical Industry” 3rd Edition, ABB Oil and
Gas, Etterstad, Oslo, 152 pp.
5. Ken, A. and Maurice, S. (1998), “Surface Production Operations”, Vol. 1, 2nd Edition,
Gulf Professional Publishing, Houston, TX, pp. 244 – 284.
6. Mauricio, G. P. (2009), “Lecture Notes on Production Engineering” African University
of Science and Technology, Abuja, Nigeria.

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 116
Advanced Production Engineering 2019/2020 Academic Year

APPENDIX A
Compressibility of low-molecular-weight natural gases

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 117
Advanced Production Engineering 2019/2020 Academic Year

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 118
Advanced Production Engineering 2019/2020 Academic Year

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 119
Advanced Production Engineering 2019/2020 Academic Year

Compiled by Dr Harrison Osei Dept. of Petroleum Engineering, UMaT, Tarkwa Page 120

You might also like