Sie sind auf Seite 1von 239

Artificial Lift Methods

Artificial Lift Overview


INTRODUCTION TO ARTIFICIAL LIFT
The Inflow Performance Relationship, or IPR, defines a wells flowing production potential:

(1)

q = PI (Pavg - Pwf)
where q = production rate, B/D
PI
= productivity index, B/D/psi
Pavg, or = average reservoir pressure, psi
Pwf
= flowing bottomhole pressure, psi

(Figure 1) illustrates this relationship for a solution gas drive reservoir. Note that for a given
average reservoir pressure and productivity index, P wf determines the wells production potential.
The lower the flowing bottomhole pressure, the higher the production rate. The wells maximum or
absolute flow potential would correspond to a Pwf of zero.

Figure 1

A well never actually attains its absolute flow potential, because in order for it to flow, P wf must
exceed the backpressure that the producing fluid exerts on the formation as it moves through the
production system. This backpressure or bottomhole pressure has the following components:

Hydrostatic pressure of the producing fluid column


Friction pressure caused by fluid movement through the tubing, wellhead and surface
equipment

Kinetic or potential losses due to diameter restrictions, pipe bends or elevation changes. In
most production systems, these are not of the same magnitude as hydrostatic or friction
pressures; in others, however, they may be significant.

Artificial lift is a means of overcoming bottomhole pressure so that a well can produce at some
desired rate, either by injecting gas into the producing fluid column to reduce its hydrostatic
pressure, or using a downhole pump to provide additional lift pressure downhole.
We tend to associate artificial lift with mature, depleted fields, where P avg has declined such that
the reservoir can no longer produce under its natural energy. But these methods are also used in
younger fields to increase production rates and improve project economics.

ARTIFICIAL LIFT METHODS


The two major categories of artificial lift are gas lift and pump-assisted lift. Plunger lift, which
combines elements of both categories, is used primarily in gas and high-GOR wells to produce
relatively small volumes of liquid.

GAS LIFT
Gas lift involves injecting high-pressure gas from the surface into the producing fluid column
through one or more subsurface valves set at predetermined depths ( Figure 2: Gas lift system.
Courtesy Weatherford International Ltd).

Figure 2

There are two main types of gas lift:

Continuous gas lift, where gas is injected in a constant, uninterrupted stream. This lowers the
overall density of the fluid column and reduces the hydrostatic component of the flowing
bottomhole pressure. Thus, for a given average reservoir pressure and productivity index,
the well is able to flow at a higher rate. This method is generally applied to wells with high
productivity indexes and high bottomhole pressures relative to their depths.
Intermittent gas lift, which is designed for lower-productivity wells. In this type of gas lift
installation, a volume of formation fluid accumulates inside the production tubing. A highpressure slug of gas is then injected below the liquid, physically displacing it to the surface.
As soon as the fluid is produced, gas injection is interrupted, and the cycle of liquid
accumulation-gas injection-liquid production is repeated.

The availability of gas and the costs for compression and injection are major considerations in
planning a gas lift installation. Where these gas injection requirements can be satisfied, gas lift
offers a flexible means of optimizing production. It can be used in deviated or crooked wellbores,
and in high-temperature environments that might adversely affect other lift methods, and it is
conducive to maximizing lift efficiency in high-GOR wells. Wireline-retrievable gas lift valves can
be pulled and reinstalled without pulling the tubing, making it relatively easy and economical to
modify the design.
On the negative side, gas lift system costs can adversely impact profitability if the source gas
requires additional processing or surface compression. Lift efficiency can be reduced by corrosion
and paraffin, which increase friction and backpressure. Tubing size and surface flowline length
also affect system efficiency. Another disadvantage of gas lift is its difficulty in fully depleting lowpressure, low-productivity wells. Also, the start-and-stop nature of intermittent gas lift may cause
downhole pressure surges and lead to increased sand production.

PUMP-ASSISTED LIFT
Downhole pumps are used to increase pressure at the bottom of the tubing string by an amount
sufficient to lift fluid to the surface. These pumps fall into two basic categories: positive
displacement pumps and dynamic displacement pumps.
A positive displacement pump works by moving fluid from a suction chamber to a discharge
chamber. The suction chamber volume increases as the discharge chamber volume decreases,
causing fluid to enter the suction chamber. As the cycle reverses, the suction volume decreases
and the discharge volume increases, forcing the fluid to the discharge end of the pump. This basic
operating principle applies to reciprocating rod pumps, hydraulic piston pumps and progressive
cavity pumps (PCPs).
A dynamic displacement pump works by causing fluid to move from inlet to outlet under its own
momentum, as is the case with a centrifugal pump. Dynamic displacement pumps commonly used
in artificial lift include electrical submersible pumps (ESPs) and hydraulic jet pumps.

Reciprocating Rod Pump Systems


A reciprocating rod pump system is made up of the following components ( Figure 3: Rod pumping
system. Courtesy Weatherford International Ltd):

Figure 3

A beam pumping unit, operated by an electric motor or gas engine.


A string of steel or fiberglass sucker rods that connect the beam pumping unit to the downhole
pump.

A subsurface pump, which consists of a barrel, plunger and valve assembly that moves fluid
through the tubing and up to the surface.

Beam pumping is the most common and arguably the most recognizable artificial lift method. It
can be used for a wide range of production rates and operating conditions, and rod pump systems
are relatively simple to operate and maintain. However, the volumetric efficiency (capacity) of a rod
pump is lower in wells with high gas-liquid ratios, small tubing diameters or deep producing
intervals. Surface equipment requires a large amount of space compared with other lift methods,
and its initial installation may involve relatively high capital costs.

Progressive Cavity Pump (PCP) Systems


A progressive cavity pump consists of a spiral rotor that turns eccentrically inside an elastomerlined stator (Figure 4: PCP system. Courtesy Weatherford International Ltd).

Figure 4

As the rotor turns, cavities between the threads of the pump rotor and stator move upward. The
rotor is most often powered by rods connected to a motor on the surface, although some
assemblies are driven by subsurface electric motors.
Progressive cavity pumps are commonly used for dewatering coalbed methane gas wells, for
production and injection applications in waterflood projects and for producing heavy or high-solids
oil. They are versatile, generally very efficient, and excellent for handling fluids with high solids
content. However, because of the torsional stresses placed on rod strings and temperature
limitations on the stator elastomers, they are not used in deeper wells.

Hydraulic Pump Systems


Hydraulic pump systems use a power fluidusually light oil or waterthat is injected from the
surface to operate a downhole pump. Multiple wells can be produced using a single surface power
fluid installation (Figure 5: Hydraulic pumping system. Courtesy Weatherford International Ltd).

Figure 5

Downhole hydraulic pumps may be either of two types.

With a reciprocating hydraulic pump, the injected power fluid operates a downhole fluid
engine, which drives a piston to pump formation fluid and spent power fluid to the surface.

A jet pump is a type of hydraulic pump with no moving parts. Power fluid is injected into the
pump body and into a small-diameter nozzle, where it becomes a low-pressure, highvelocity jet. Formation fluid mixes with the power fluid, and then passes into an expandingdiameter diffuser. This reduces the velocity of the fluid mixture, while causing its pressure to
increase to a level that is sufficient to lift it to the surface

Hydraulic pumps can be used at depths from 1000 to 17,000 feet and are capable of producing at
rates from 100 to 10,000 B/D. They can be hydraulically circulated in and out of the well, thus
eliminating the need for wireline or rig operations to replace pumps and making this system
adaptable to changing field conditions. Another advantage is that heavy, viscous fluids are easier
to lift after mixing with the lighter power fluid.
Disadvantages of hydraulic pump systems include the potential fire hazards if oil is used as a
power fluid, the difficulty in pumping produced fluids with high solids content, the effects of gas on
pump efficiency and the need for dual strings of tubing on some installations.

Electrical Submersible Pumps


An electric submersible pumping (ESP) assembly consists of a downhole centrifugal pump driven
by a submersible electric motor, which is connected to a power source at the surface ( Figure 6:
ESP system. Courtesy Weatherford International Ltd).

Figure 6

The pump and motor assembly, which may be several hundred feet long, is connected to the
surface by an armored cable that provides electric power and control. On a cost-per-barrel basis,
ESP systems are among the most efficient and economical of lift methods. Fluid volumes ranging
from 100 to 60,000 B/D, including high water-cut fluids, can be handled by ESP systems. These
systems can be installed in high-temperature wells (above 350F) using high-temperature motors
and cables. The pumps can be modified to lift corrosive fluids and sand. ESP systems can be
used in high-angle and horizontal wells if placed in straight or vertical sections of the well.
ESP pumps can be damaged from gas lock. In wells producing high GOR fluids, a downhole gas
separator must be installed. Another disadvantage is that ESP pumps have limited production
ranges determined by the number and type of pump stages; changing production rates requires
either a pump change or installation of a variable-speed surface drive. The tubing must be pulled
for pump repairs or replacement.

PLUNGER LIFT
Plunger lift is the only artificial lift method that relies solely on the wells natural energy to lift fluids.
The plunger, traveling inside the tubing, moves upward when the pressure of the gas below it is
greater than the pressure of the liquid above it ( Figure 7: Plunger lift system. Courtesy
Weatherford International Ltd ) .

Figure 7

As the plunger travels to the surface, it creates a solid interface between the lifted gas below and
produced fluid above to maximize lifting energy. Any gas that bypasses the plunger during the
lifting cycle flows up the production tubing and sweeps the area to minimize liquid fallback.
Plunger lift provides a cost-effective method of artificial lift that can be used to efficiently produce
both gas wells with fluid loads and high GOR oil wells.

SELECTING AN ARTIFICIAL LIFT METHOD


Artificial lift considerations should ideally be part of the well planning process. Future lift
requirements will be based on the overall reservoir exploitation strategy, and will have a strong
impact on the well design.

INITIAL SCREENING CRITERIA


Tables 1 and 2 below summarize some of the key factors that influence the selection of an artificial
lift method.
Table 1
Reservoir and Hole Considerations in Selecting an Artificial Lift Method
(after Brown, 1980)
Reservoir Characteristics:
IPR

A wells inflow performance relationship defines its production


potential

Liquid production rate

The anticipated production rate is a controlling factor in selecting


a lift method; positive displacement pumps are generally limited to
rates of 4000-6000 B/D.

Water cut

High water cuts require a lift method that can move large volumes
of fluid

Table 1
Reservoir and Hole Considerations in Selecting an Artificial Lift Method
(after Brown, 1980)
Reservoir Characteristics:
Gas-liquid ratio

A high GLR generally lowers the efficiency of pump-assisted lift

Viscosity

Viscosities less than 10 cp are generally not a factor in selecting a


lift method; high-viscosity fluids can cause difficulty, particularly in
sucker rod pumping

Formation volume
factor

Ratio of reservoir volume to surface volume determines how


much total fluid must be lifted to achieve the desired surface
production rate

Reservoir drive
mechanism

Depletion drive reservoirs: Late-stage production may require


pumping to produce low fluid volumes or injected water.
Water drive reservoirs : High water cuts may cause problems for
lifting systems
Gas cap drive reservoirs : Increasing gas-liquid ratios may affect
lift efficiency.

Other reservoir
problems

Sand, paraffin, or scale can cause plugging and/or abrasion.


Presence of H2S, CO2 or salt water can cause corrosion.
Downhole emulsions can increase backpressure and reduce
lifting efficiency. High bottomhole temperatures can affect
downhole equipment.

Hole Characteristics:
Well depth

The well depth dictates how much surface energy is needed


to move fluids to surface, and may place limits on sucker
rods and other equipment.

Completion type

Completion and perforation skin factors affect inflow


performance.

Casing and tubing


sizes

Small-diameter casing limits the production tubing size and


constrains multiple options. Small-diameter tubing will limit
production rates, but larger tubing may allow excessive fluid
fallback.

Wellbore deviation

Highly deviated wells may limit applications of beam


pumping or PCP systems because of drag, compressive
forces and potential for rod and tubing wear.

Table 2
Surface and Field Operating Considerations in Selecting an Artificial Lift Method
(after Brown, 1980)
Surface Characteristics:
Flow rates

Flow rates are governed by wellhead pressures and


backpressures in surface production equipment (i.e.,
separators, chokes and flowlines).

Table 2
Surface and Field Operating Considerations in Selecting an Artificial Lift Method
(after Brown, 1980)
Surface Characteristics:
Flowline size and length

Flowline length and diameter determines wellhead pressure


requirements and affects the overall performance of the
production system.

Fluid contaminants

Scale, paraffin or salt can increase the backpressure on a


well.

Power sources

The availability of electricity or natural gas governs the type


of artificial lift selected. Diesel, propane or other sources
may also be considered.

Field location

In offshore fields, the availability of platform space and


placement of directional wells are primary considerations. In
onshore fields, such factors as noise limits, safety,
environmental, pollution concerns, surface access and well
spacing must be considered.

Climate and Physical


environment

Affect the performance of surface equipment.

Field Operating Characteristics:


Long-range recovery plans

Field conditions may change over time.

Pressure maintenance
operations

Water or gas injection may change the artificial lift


requirements for a field.

Enhanced oil recovery


projects

EOR processes may change fluid properties and require


changes in the artificial lift system.

Field automation

If the surface control equipment will be electrically powered,


an electrically powered artificial lift system should be
considered.

Availability of operating and


service personnel and
support services

Some artificial lift systems are relatively low-maintenance;


others require regular monitoring and adjustment. Servicing
requirements (e.g., workover rig versus wireline unit) should
be considered. Familiarity of field personnel with equipment
should also be taken into account.

Clegg, Bucaram and Hein (1993), in a piece written for the SPE Distinguished Author Series,
observe that selecting the proper artificial lift method is critical to the long-term profitability of most
producing oil and gas wells. They list 31attributes for comparing the eight most common artificial
lift techniques (continuous and intermittent gas lift, beam pumping, progressing cavity pumping,
hydraulic pumping, electric submersible pumping, jet pumping and plunger lift), and provide
practical guidelines for assessing each methods capabilities. These are summarized as follows:
Design considerations and overall comparisons:

Capital cost
Downhole equipment

Operating costs
Reliability

Efficiency
Flexibility

Salvage value
System (total)

Miscellaneous [operating]

Usage/outlook

problems
Normal operating considerations:

Prime mover flexibility


Surveillance
Testing

Casing size limits


Depth limits
Intake capabilities
Noise level

Time cycle and pump-off controllers application


Obtrusiveness
Artificial lift considerations:

Corrosive/scale handling ability


Crooked/deviated holes
Multiple completions
Gas-handling ability
Offshore application

Paraffin-handling capability

Slim-hole completions
Solids/sand-handling ability
Temperature limitations
High-viscosity fluid handling
High-volume lift capabilities

Low-volume lift capabilities

Finally, Table 3 (from Weatherford International Ltd., 2005) summarizes typical characteristics and
applications for each form of artificial lift. These are general guidelines, which vary among
manufacturers and researchers. Each application needs to be evaluated on a well-by-well basis.
Table 3: Artificial Lift MethodsCharacteristics and Areas of Application
(after Weatherford, 2005)
Operati
ng
Parame
ters

Positive displacement
pumps

Dynamic
displacement
pumps
ESP

Gas lift

Plunge
r lift

5000 to
10000 ft

To
8000 ft

Rod
pump

PCP

Hydrau
lic
Piston

Hydrau
lic Jet

Typical
Operati
ng
Depth
(TVD)

100 to
11000
ft

2000 to
4500 ft

7500 to

5000 to

10000
ft

10000
ft

Maximu
m
Operati
ng
Depth
(TVD)

16000
ft

6000 ft

17000
ft

15000
ft

15000
ft

15000 ft

20000
ft

Typical
Operati
ng
Volume

5 to
1500
BFPD

5 to
2200
BFPD

50 500
BFPD

100 to
30000
BFPD

300 4000
BFPD

100 10000
BFPD

1-5
BFPD

Maximu

6000

4500

4000

40000

>1500

30000

200

Table 3: Artificial Lift MethodsCharacteristics and Areas of Application


(after Weatherford, 2005)
Operati
ng
Parame
ters

Positive displacement
pumps

Dynamic
displacement
pumps

Gas lift

Plunge
r lift

Rod
pump

PCP

Hydrau
lic
Piston

ESP

Hydrau
lic Jet

m
Operati
ng
Volume

BFPD

BFPD

BFPD

BFPD

0
BFPD

BFPD

BFPD

Typical
Operati
ng
Temper
ature

100 350 F

75 150 F

100 250 F

100 250 F

100 - 250
F

120 F

[40-177
C]

[24-65
C]

[40120
C]

[40120
C]

[40-120
C]

Maximu
m
Operati
ng
tempera
ture

550 F

250 F

500 F

400 F

500 F

400 F

500 F

[288
C]

[120
C]

[260
C]

[205
C]

[260
C]

[205 C]

[260
C]

Typical
Wellbor
e
Deviatio
n

0 - 20
deg
landed
pump

N/A

0 - 20
deg
landed
pump

0 - 20
deg
hole
angle

0 - 50
deg

N/A

Maximu
m
Wellbor
e
Deviatio
n

0 - 90
deg
landed
pump

0 - 90
deg

0 - 90
deg

0 - 90
deg

80 deg

< 15
deg/10
0 ft

< 15
deg/10
0 ft

70 deg,
short to
medium
radius

Corrosio
n
handling

Good
to
Excelle
nt

Fair

Good

Good

Excelle
nt

Good to
excellent

Excelle
nt

Gas
handling

Fair to
good

Good

Fair

Fair

Good

Excellent

Excelle
nt

Solids
handling

Fair to
good

Excelle
nt

Poor

Fair

Good

Good

Poor to
Fair

>8
API

< 35
API

>8
API

> 10
API

>8
API

> 15
API

GLR =
300
SCF/B
bl per
1000 ft
of

Fluid
gravity

0 - 90
deg

< 24
deg/10
0 ft

[50 C]

Table 3: Artificial Lift MethodsCharacteristics and Areas of Application


(after Weatherford, 2005)
Operati
ng
Parame
ters

Positive displacement
pumps
Rod
pump

PCP

Hydrau
lic
Piston

Dynamic
displacement
pumps
ESP

Gas lift

Plunge
r lift

Hydrau
lic Jet
depth

Servicin
g

Workov
er or
pulling
rig

Workov
er or
pulling
rig

Hydrau
lic or
wirelin
e

Workov
er or
pulling
rig

Hydrau
lic or
wirelin
e

Wireline
or
workover
rig

Wellhe
ad
catcher
or
wireline

Prime
mover

Gas or
electric

Gas or
electric

Multicylinde
r or
electric

Electric
motor

Multicylinde
r or
electric

Compres
sor

Wells
natural
energy

Offshor
e
applicati
ons

Limited

Good

Good

Excelle
nt

Excelle
nt

Excellent

N/A

System
efficienc
y

45% 60%

40% 70%

45% 55%

35% 60%

10% 30%

10% 30%

N/A

ECONOMICS OF ARTIFICIAL LIFT


The features, benefits and limitations of one artificial lift method are relative to those of the other
methods under consideration. Each method should be evaluated from the standpoint of
comparative economics. Brown (1980) lists six critical bases of comparison:

Initial capital cost


Monthly operating expense
Equipment life
Number of wells to be lifted
Surplus equipment availability
Expected producing life of well(s)

Capital cost considerations may favor one type of system over another, particularly when there is
significant uncertainty regarding well performance characteristics or reserve volumes. Gas lift is
not likely to be a good option for a one or two-well system, for exampleparticularly if it requires
adding surface compression facilities. For multiple wells, however, it may be a very economical
choice. Hydraulic pumping is likewise less costly when multiple wells are operated from a central
injection facility.
Projected operating costs also figure into the selection of an artificial lift method. High gas prices
will reduce the profitability of gas lift, particularly if it becomes necessary to purchase additional
gas for injection. But gas lift may be an attractive option in a remote field where there is no market

for produced gas. In the same way, in places where electricity is not readily available, submersible
pumps will be less attractive compared to gas lift or other forms of pump-assisted lift.
System reliability and easy access to repair equipment and services must also be considered.
Sometimes, the prevalence of a particular type of lift equipment in a given area will make that
system more attractive.
If a well is expected to have a short producing life, capital and operating costs will play an
important role in the overall field economics and will affect the choice of an artificial lift system.
It is clear that for each well or field situation, a number of factors will affect the choice of artificial lift
system. Equipment manufacturers can explain important advantages and disadvantages of
different systems. Each type of artificial lift method has economic and operating limitations that
can make it more or less desirable when compared to others. Similarly, one artificial lift system will
usually have at least one advantage over all others for a given set of operating conditions.

Gas Lift
GAS LIFT SYSTEM OVERVIEW
Gas lift is a four-step process (Figure 1: Gas lift system):

Figure 1

1. Natural gas is compressed at the surface and routed to individual wells.


2. This lift gas is injected downhole and into the produced fluid stream through one or more
valves set at specified depths (most commonly, the gas is injected into the production tubing
from the casing-tubing annulus).
3. The lift gas and formation fluids are produced to the surface.
4. The gas and liquids are separated; the gas is then treated and sent either to compression or
to sales.

In most wells, gas is injected continuously into the produced fluid stream. This continuous gas lift
process reduces the backpressure on the formation by reducing the densityand therefore the
hydrostatic pressureof the produced fluid (Figure 2: Continuous gas lift).

Figure 2

Continuous gas lift is typically used in higher productivity wells to handle rates ranging from 100
up to 30,000 B/D. In wells with very high productivity indexes, even higher rates can be attained by
injecting gas into the tubing and producing fluids through the casing-tubing annulus.
Intermittent gas lift employs much of the same equipment as continuous lift, but its operating
principle is completely different. Rather than lowering the density of the produced fluid so that it
can produce in a continuous flow stream, intermittent lift works by physically displacing slugs of
liquid to the surface (Figure 3: Intermittent gas lift):

Figure 3

When a certain volume of fluid accumulates in the wellbore, gas is injected into the tubing,
where it lifts the column of fluid to the surface as a slug.
As each liquid slug is produced, gas injection is interrupted to allow the fluid volume to build
up again. Intermittent injection uses a timer or an adjustable choke located on the surface to
control the gas injection. Cycling of gas injection is regulated to coincide with the
accumulation of wellbore fluids.

Intermittent lift is generally used in wells with limited inflow potential (i.e., high productivity index
with low average reservoir pressure or, alternatively, low productivity index with high reservoir
pressure).

As is true for other artificial lift methods, gas lift offers a number of benefits, and at the same time
has some inherent limitations (Brown, 1982; Tak cs, 2005). Advantages include:

Flexibility in handling a wide range of production rates; can convert from continuous to
intermittent lift as reservoir pressure or well productivity declines.

Relatively good solids-handling capabilities.


Suitability for producing high-GLR wells (unlike pump-assisted lift, where gas production is
usually detrimental to system efficiency).

Can be used in deviated wells


Installations can be designed for servicing with wireline units; gas lift valves can be run and
retrieved without having to pull tubing string.

Relatively low-profile surface wellhead equipment, takes up little surface space.


Can easily manage high bottomhole temperatures or corrosive environments
Most installations provide full-bore tubing strings, which facilitate downhole surveys, well
monitoring and workover.

Some of the main limitations and disadvantages of gas lift systems include:

Obtaining sufficient amounts of lift gas.


The need to provide compression and gas treatment facilities.
Generally lower energy efficiency than other lift methods.
Cannot reduce bottomhole pressure to the low levels attainable by pump-assisted lift.

DOWNHOLE INSTALLATIONS
The key subsurface components of a gas lift system are the gas lift valves that regulate the flow
of injected gas into the producing fluid column. These pressure-operated devicesusually 1 or 1.5
inches in diameter and about 16 to 24 inches longare placed in mandrels that are set at
selected depths in the tubing string, most often in a conventional or side pocket configuration
(Figure 1: Conventional and side pocket mandrel installations).

Figure 1

In a conventional or tubing retrievable valve-mandrel configuration, the valves are run with the
tubing string, and the tubing must be pulled in order to repair or replace a valve. (Figure 2)
shows a valve designed for a conventional gas lift installation. (Camco conventional
injection-pressure operated gas lift valves, Types J50 and J40. Courtesy of Schlumberger.)

Figure 2

Side pocket mandrels are designed to accommodate valves in parallel with the tubing
(Figure 3: Camco KBMM Series side pocket mandrel. Courtesy of Schlumberger.). These
are the most common types of mandrels in use, their main attraction being that they enable
gas lift valves to be run and retrieved on wireline. This eliminates the need to pull the tubing
for valve repairs or adjustments.

Figure 3

INSTALLATION TYPES
The type of downhole installation used in a gas lift well depends on whether it is to be placed on
continuous or intermittent lift. This in turn depends on its present and anticipated future inflow
performance. Other considerations include completion type, casing diameters and wellbore
deviation. As is true for any type of downhole installation, a gas lift system should be designed
with enough flexibility to minimize the number of workovers required over a wells producing life.
Except for the annular flow installation mentioned at the end of this section, all of the downhole
installations discussed here are designed for gas injection through the casing and production
though the tubing string.

OPEN INSTALLATION
An open gas lift installation is one in which the tubing string is suspended in the well without a
packer, and the casing and tubing are in communication. This, the oldest type of gas lift
installation, has several major disadvantages:

Only a fluid seal in the annulus prevents gas from blowing around the bottom of the tubing.
This results in wasted gas, additional backpressure on the formation, and reduced
production rates.

Without a packer, the lower gas lift valves may be submerged in well fluids because the fluid
rises in the annulus every time the well is shut-in. This may lead to valve corrosion.

When production resumes, the fluid must flow back through the gas lift valves, causing the
valves to wear out faster.

The open installation is not normally recommended, and it is used nowadays only when a packer
cannot be installed.

SEMI-CLOSED INSTALLATION
A semi-closed installation has a packer installed in the tubing to seal off the tubing-casing annulus,
as shown in the continuous gas lift well of (Figure 4) (Semi-closed gas lift installation).

Figure 4

This is the most common type of installation for continuous gas lift wells. It eliminates most of the
characteristic disadvantages of the open installationthe packer keeps produced fluids from
entering the annulus, and prevents the casing pressure from directly communicating with the
formation.
This installation is also used in intermittent gas lift wells. It is not the best choice for wells
exhibiting very low bottomhole pressures, however, because it is possible that gas injected into the
tubing string may place additional backpressure on the formation when the operating valve is
open.

CLOSED INSTALLATION
A closed installation (Figure 5) is similar to a semi-closed installation, except that a standing valve
is placed in the tubing string below the bottom gas lift valve to prevent fluids from moving
downward. Thus, high-pressure gas injected into the tubing from the annulus cannot increase
backpressure on the formation, and any produced fluids standing in the tubing will not flow back
into the formation. These features make the closed installation the option of choice for intermittent
gas lift.

Figure 5

CHAMBER INSTALLATIONS
A chamber installation can greatly increase production rates, especially in wells with low
bottomhole pressures and high productivity indexes. It is used in intermittent lift operations to
increase the volume of fluids in the wellbore prior to lifting, without significantly increasing the
backpressure on the formation.
One type of chamber installation consists of a lower packer and an upper, or bypass packer. This
two-packer chamber installation works as follows:

As the chamber fills with fluid, gas in the chamber passes through a bleed valve into the
tubing.

When the chamber is filled, a slug of gas is injected down the annulus to open the operating
valve.

The gas in the chamber forces the fluid to enter the tubing through a perforated nipple above
the bottom packer (Figure 6: Chamber installationfluid entry).

Figure 6

When all the fluid in the chamber above the nipple is forced into the tubing, gas follows behind
the slug and forces it to the surface (Figure 7: Chamber installationfluid displacement).

Figure 7

The operating valve should close when the slug reaches the surface, at which time the filling
cycle begins again.

In wells where it is not feasible to set two packers, an insert chamber installation may be used, in
which the chamber is formed by a larger-diameter section of pipe at the bottom of the tubing
string. The principles of operation are the same as for the two-packer installation. Alternative
configurations may be designed for special circumstances.

SLIM HOLE INSTALLATIONS

Slim hole completions are often used in lower productivity wells. These normally use a string of 23/8 to 3-1/2-inch OD pipe as the production casing. Smaller size tubing, (e.g., 1 to 1-1/2-inches in
diameter) is then run inside this casing (Figure 8: Slim hole gas lift completion).

Figure 8

Slim hole completions are especially useful for producing from two or more zones without
commingling. Production rates for continuous gas lift will depend on the ID of the tubing, but can
range from 150 B/D for 3/4-inch tubing to 900 B/D for 1 1/2-inch ID tubing. The rates for
intermittent injection in slim hole wells are considerably lower.

DUAL INSTALLATIONS
Gas lift wells, like flowing wells, can be designed with parallel or concentric dual tubing strings
(Figure 9: Dual gas lift installations). The two most common configurations are (1) parallel strings
of 2 3/8-inch OD tubing inside 7-inch casing and (2) parallel strings of 3 1/2-inch OD tubing inside
9 5/8-inch casing (API RP 11V8, 2003).

Figure 9

Gas is supplied through the tubing-casing annulus, and injected into two separate tubing strings.
The zones may be produced using the same type of lift method (i.e., both on continuous or both
on intermittent lift), or the two methods may be used in combination (e.g., one zone is on
intermittent lift while the other is on continuous lift). For a dual installation to work, the valves must
be spaced to prevent interference between the two zones, and selected so that the desired
amounts of gas are injected for each zone.

COILED TUBING INSTALLATIONS


Coiled tubing can be used to convert a flowing well to gas lift without having to pull the main
production tubing string, simply by placing conventional gas lift mandrels at the appropriate depths
on the coiled tubing string and then running it inside the production tubing. The smaller diameter of
the coiled tubing restricts its use to lower-productivity wells.

ANNULAR FLOW INSTALLATIONS


In most gas lift operations, production is confined to the tubing because of safety issues,
regulatory requirements and company operating policies. This is especially true offshore. In some
areas, howeversuch as the Middle East, where wells produce at rates up to 80,000 B/D
operators use annular flow gas lift.
In this process, gas is injected down the tubing and production is from the annulus. A bull plug is
placed at the bottom of the tubing to contain the injected gas. A small-bore orifice or check valve
can also be used for this purpose.

Apart from regulatory prohibitions, annular flow installations are limited in their application due to
concerns about casing corrosion, high injection requirements and the potential for fluid slugging as
production rates decline.

GAS LIFT VALVES


Depending on where it is placed in the tubing string, a gas lift valve may be used as an operating
valve or an unloading valve.

The operating valve is the deepest valve in the string. It is designed and placed to ensure that
the proper amount of gas is injected into the tubing at the appropriate depth to optimize well
performance.
Unloading valves are set are predetermined depths above the operating valve. They are used
to progressively reduce the static fluid level in a well when it is first brought on production (or
returned to production after having been killed for workover or other operations). As gas is
injected into the well, each valve opens in sequence from top to bottom. The gas displaces
the liquid in the well, u-tubing it through the unloading valve and displacing it to the
surface. As the liquid level drops below the valve, that valve closes and the valve below it
opens, and so on down to the operating valve. The unloading valves remain closed during
normal production.

Individual valve specifications will depend on whether the well is being produced by continuous or
intermittent lift.

VALVE TYPES
Gas lift valves are pressure-operated valves, so-called because they are designed to open or
close in response to gas injection pressure, fluid production pressure or both. The pressure that
controls the valve operation is that which is exposed to the largest area in the valve.
The most common types of gas lift valves are described below:

Injection pressure-operated (IPO) valve: Increase in gas injection pressure opens valve;
decrease in gas injection pressure closes valve (Figure 1: Camco retrievable injection
pressure operated gas lift valves, BK series and R-20-02 series. Courtesy of Schlumberger.)

Figure 1

Production-pressure-operated (PPO) and Differential-pressure-operated valves :


Increase in production pressure opens valve; decrease in production pressure closes valve
(Figure 2: Camco retrievable production pressure operated gas lift valve, type BKF-12.
Courtesy of Schlumberger).

Figure 2

Throttling valve: Increase in gas injection pressure opens valve; decrease in gas injection
pressure or production pressure closes valve.

Combination valves: Increase in production pressure opens valve; decrease in gas injection
pressure or production pressure closes valve.

Where injection takes place in the annulus and production takes place through the tubing, the gas
injection pressure is commonly referred to as the casing pressure and the production pressure is
referred to as the tubing pressure.

INJECTION PRESSURE-OPERATED VALVES


An injection-pressure-operated (IPO) valve opens and closes in response to changes in gas
injection pressure. One type of IPO valve, represented schematically in ( Figure 3), operates as
follows:

Figure 3

The dome is charged to a specified pressure with nitrogen, at a controlled surface


temperature.

The bellows serve as a flexible or responsive element. Movement of the bellows causes the
valve stem to rise and fall, and the ball to open and close over the port.

When the port is open, the annulus and tubing are in communication
Because the area of the bellows is much larger than the area of the port, and since the
bellows is exposed to casing pressure, it is casing pressure that controls the operation of the
valve.
A buildup in injection pressure opens the valve, and a reduction in injection pressure closes it

IPO valves may be classified as either unbalanced or balanced, depending on whether the
production pressure plays a role in opening the valve.

A balanced valve opens only at a specified pressure and does not respond to pressure
changes in the tubing. By design, tubing pressure cannot act on the port or bellows; thus,
the valve is not affected by changes in tubing pressure. Only casing pressure can move the
bellows and control the flow of gas from the casing into the tubing.
An unbalanced valve is one in which the (1) opening or (2) opening and closing pressure are
affected by the production pressure. The valve is kept closed by a nitrogen-charged dome
that loads the bellows. The bellows is attached to a stem that moves a ball and controls gas
flow into the tubing. When tubing pressure is high enough, the ball moves up and the valve
opens.

The valve in (Figure 1) is unbalanced, since the tubing pressure can open the port. It may also be
classified as a single-element valve, since the charge pressure in the dome represents the only
control on the valves operation. In contrast, a double-element valve has two loading elements: the
pressure-charged dome and a spring. The spring provides a closing force, which ensures that if
the bellows is ruptured, the valve can still close when needed. Double element valves can be used
in both continuous and intermittent flow gas lift installations.

PRODUCTION PRESSURE-OPERATED VALVES


In a production pressure-operated (PPO) valve, the port is exposed to the injection pressure and
the bellows is exposed to the production pressure (Figure 4: Production pressure-operated gas lift
valve ). Therefore, it is the production pressure that controls the operation of the valve.

Figure 4

PPO valves are double element valves, having both a spring and dome (that may or may not be
charged) to supply the valve closing force. Most manufacturers of this valve type charge the dome
only when high valve-setting pressures require a supplement to the spring force.
Another type of production pressure-operated valve is called a differential-pressure-operated
valve. This valve opens and closes in response to tubing pressure relative to the casing pressure.
It does not have a pressure-charged dome, but has only a spring acting against the stem and ball.

THROTTLING VALVES
An IPO valve closes only when the casing pressure falls below the dome pressure. If a tapered
seat is used, however, then the valves closing becomes somewhat sensitive to tubing pressure

(Figure 5: Throttling gas lift valve) . The tapered seat allows the port area to sense the tubing
pressure when the valve is open. This type of valve, called a throttling valve, responds to both
tubing and casing pressure, even when it is open.

Figure 5

If the tubing pressure is lower than the casing pressure, the throttling valve can close even before
the casing pressure has dropped to the dome pressure. In fact, a throttling valve will close with a
reduction in tubing pressure - even though the casing pressure is held constant. Throttling valves
are also known as proportional valves or continuous flow valves.
A throttling valve, then, requires a buildup in tubing or casing pressure to open, and a reduction in
tubing or casing pressure to close. During continuous gas lift, the injection gas pressure is held
constant by a regulator at the surface; as a result, the valve opens and closes only in response to
changes in tubing pressure.

PILOT VALVES
A pilot valve is a variable-spread valve that is used to unload or kick off a well. It has both a small
port to control the spread (i.e., the difference between the opening and closing pressures), and a
larger port used for more efficient gas flow. This type of valve is often used for intermittent gas lift
operations, which benefit from a valve with a large port size while also keeping close control over
the spread (Figure 6: Pilot Valve Operation) .

Figure 6

DOME CHARGE PRESSURE CORRECTIONS


A pressure-operated gas lift valve is charged to its specified dome pressure under controlled
surface temperature conditions. Its use in a gas lift installation, however, is based on the charge
pressure at its setting depth. The surface pressure therefore has to be corrected for changes in
temperature and the gas compressibility factor.
Brown (Vol. 2a, 1980), Winkler and Smith (1962), Takcs (2005) and others have published charts
that can be used to determine the nitrogen dome charge pressure at a given downhole
temperature. Winkler and Eads (1989) present the following formulas, which incorporate the
effects of the deviation factor for nitrogen:

For Pd < 1238 psi, where Pd = dome charge pressure at 60 F, psia:


Pd = Pd + (-0.00226 + 0.001934 Pd + 3.054 x 10-7 (Pd)2) x (Tvalve - 60)
For Pd > 1238 psi:
Pd = Pd + (-0.267 + 0.002298 Pd + 1.84 x 10-7 (Pd)2) x (Tvalve - 60)

Example:
A gas lift valve has a dome charge pressure of 500 psia at a surface temperature of 60 F. What is
this pressure at the valve setting depth of 6000 ft, where the temperature is equal to 144 F?
Solution:
Pd = 500 psia; Pd < 1238 psi
Pd = Pd + (-0.00226 + 0.001934 Pd + 3.054 x 10-7 (Pd)2) x (Tvalve - 60)
Pd = 500+ (-0.00226 + 0.001934 (500) + 3.054 x 10 -7 (500)2) x (144 - 60)
Pd = 587 psi

GAS LIFT VALVE SELECTION

The type of gas lift valve used in a given installation will depend on whether the well will be placed
on continuous or intermittent lift. Some types of valves are suitable for either continuous or
intermittent injection; these may be worth considering if the lift method has not been determined or
if well performance is marginal.

CONTINUOUS GAS LIFT


The ideal continuous lift installation would be one in which gas is injected in a steady stream,
resulting in more-or-less constant pressures and production rates according to the original design.
In reality, production rates and pressures fluctuate from the design parameters, often on a day-today basis. The operating valve thus needs to be sensitive to production pressure when it is in the
open position. As production pressure decreases, (i.e., tubing pressure for a tubing installation)
the valve should begin to throttle closed to decrease gas throughput; as it increases, the valve
should open to increase gas throughput. This proportional response maintains the established
flowing wellhead pressure and keeps a constant pressure inside the tubing. Therefore, the best
type of operating valve to use for continuous gas flow would be a throttling valve or one with
identical gas throughput characteristics.

INTERMITTENT GAS LIFT


Operating valves used for intermittent lift must be designed for immediate or snap opening and
closing. Immediate opening ensures that the gas is injected as a compressed slug and not as a
gradual stream that bubbles up through the liquid, while quick valve closure prevents excess gas
from being injected. Because the operating valve must be able to handle a relatively large volume
of gas in a short time, it has to be able to open to a large port (normally between 3/8 and 3/4 inch),
and remain fully open until closing. At the same time, the valve should have a small enough
spread so that the volume of gas injected can be tightly controlled. In a single-point injection
system, these requirements can be met by using a pilot valve.
In a multipoint system, a number of valves are installed at various depths in the tubing string. At
each depth, the valve should allow sufficient gas to enter the tubing so that the fluid slug can be
moved upward to the next higher valve. As the fluid slug passes by each valve, the pressure under
the slug opens that valve, allowing more gas to be injected and supplementing the gas that has
entered through the deeper valves. The valves normally remain open until the slug is produced to
the surface.

GAS LIFT VALVE MECHANICS


The opening and closing characteristics of gas lift valves are important considerations in system
design and operation. It is important, therefore, to know how and when a valve opens and closes,
and to understand the significance of its spread, or difference between opening and closing
pressures.

PRESSURES TO OPERATE A VALVE


Consider an unbalanced IPO single-element valve in the closed position, represented
schematically in (Figure 1).

Figure 1

To calculate the valves opening and closing pressures, it is necessary to solve a force-balance
equation.

The force tending to close the valve, Fclose, is

Fclose = P d A b
Where P d is the dome charge pressure, and
A b is the bellows area.
The force tending to open the valve, Fopen, equals
Fopen = Pinj (Ab - Ap) + Pprod Ap
where Pinj is the injection pressure,
Ap is the area of the stem or port; and
Pprod is the production pressure.

(1)

(2)

If we set Fclose = Fopen, and define R = (Ap/Ab), we obtain an expression for the injection pressure
required to just open the valve at the valve setting depth, or (Pinj)open:

(3)
Example:
Determine the casing (injection) pressure required to open an IPO valve under the following
conditions: Bellows area = 0.77 square inches; Port area = 0.129 square inches; Bellows pressure
(corrected to valve depth) = 500 psi; Tubing pressure = 425 psi.
Solution:

R = A p/A b = 0.129/0.77 = 0.167

The casing pressure required to open the valve is 515 psi, or 15 psi above the bellows pressure. A
higher casing pressure is required because of the effect of the lower tubing pressure on the port
area.
We may also calculate the pressure required to close the valve once it is opened. Once again,
equate the forces tending to keep the valve opened with those tending to close it. The force
tending to close the valve is equal to:
Fclose = P d A b

(1)

The force tending to hold the valve open is equal to:


Fopen = P inj(A b - A p) + P inj A p

(4)

Note that when the valve is open, the injection pressure has replaced the production pressure
from Equation 2.
Equating these two terms yields the casing pressure required to close the valve:
(Pinj)close = P d

(5)

In this example, the bellows pressure is 500 psi, and so the valve will close at this pressure.

VALVE SPREAD
The difference between the opening pressure and the closing pressure is the spread.
Spread = (Pinj)open - (Pinj)close. )

(6)

Thus, in the preceding example, the valve spread is (515 - 500) = 15 psi.
An inspection of Equations 3 and 5 shows that the spread is a function of the ratio R, the bellows
pressure, and the tubing pressure:

Spread =

(7)

For specific bellows and tubing pressures, and since R = (A p/Ab), reducing the area of the port will
result in lowering the spread.
Spread is a particularly important consideration in intermittent lift, because it controls the volume
of gas used in each injection cycle. As the spread needed to close the operating valve increases,
the amount of gas injected during the cycle also increasesthis would call for a smaller spread in
order to reduce gas volume requirements, and therefore a smaller port size. On the other hand, a
smaller port size increases the compression horsepower requirements. Thus, there needs to be a
balance between the need to conserve gas and the need to minimize power costs.

TEST RACK OPENING PRESSURE (TRO)


When unbalanced gas lift valves are tested at the surface, they are set in special valve testers,
usually at a production pressure of zero. The injection pressure required to open the valve under
these conditions is referred to as the test rack opening, or TRO pressure:

(7)
where Pd is the dome charge pressure at surface conditions.

PRODUCTION PRESSURE EFFECT (PPE)


The Production Pressure Effect (PPE) describes the contribution that the production pressure
makes to the valves opening injection pressure, and is equal to the difference between the
opening pressure at Pprod = 0 and its actual opening pressure:

(8)
For a given valve size and port diameter, R is constant, and so PPE increases linearly with
production pressure. The slope of this line is known as the Production Pressure Effect Factor
(PPEF) and is equal to R/(1-R), or

(9)

Thus, as valve port sizes increase for a given bellows area, so does the PPEF.

FLOW CHARACTERISTICS OF A GAS LIFT VALVE


A plot of flow rate versus tubing pressure provides insight into the flow characteristics of a gas lift
valve. (Figure 2) , for example, illustrates the performance characteristics of a throttling valve
(note that in this plot, the vertical axis represents flow rate and the horizontal axis represents
tubing pressure ).

Figure 2

At very low tubing pressure, to the left of point 1, the valve is closed.
As the tubing pressure reaches point 1, the valve begins to open and gas flows from the
casing to the tubing. The flow rate increases as the port continues to open.

Throttling occurs from point 2 to point 3, at which time the port is fully opened and throttling
ends.

The maximum flow rate occurs at point 4.


As the tubing pressure increases from point 4 to point 5, the tubing and casing pressures
become balanced and the flow rate drops to zero.
During the reverse cycle (i.e., as the tubing pressure decreases) the valve opens at point 5,
throttling takes place between points 3 and 2, and the valve throttle closes between points 2
and 1.

CONTINUOUS GAS LIFT


In continuous gas lift, an uninterrupted stream of high-pressure gas is injected downhole to reduce
the flowing bottomhole pressure of the producing fluid. Injection takes place at set rates and at
pre-determined depths to optimize well performance and maximize system efficiency. In most
wells, gas is injected down the casing-tubing annulus and into the production tubing.

Continuous gas lift works as an artificial lift method because it reduces the hydrostatic component
of the bottomhole pressure. As a simple illustration, consider three shut-in wells:
Well 1

Well 2

Well 3

6000 ft

6000 ft

6000 ft

2100 psi

2100 psi

2100 psi

Fluid in tubing (TVD to surface):

Salt water

Oil

Gas

Average fluid gradient:

0.465 psi/ft

0.346 psi/ft

0.069
psi/ft

2790 psi

2076 psi

414 psi

True vertical depth (TVD):


Average reservoir pressure:

Bottomhole pressure at TVD:

Note how much lower the bottomhole pressure is in Well 3, and remember the equation that
describes the Inflow Performance Relationship:
q = PI (Pavg - Pwf)

(1)

where q = flow rate, PI = productivity index, Pavg = average reservoir pressure and Pwf = flowing
bottomhole pressure.
Clearly, if we have a well filled with liquid and can mix the liquid column with gas, we can
significantly reduce Pwf and, for a given reservoir pressure, increase the inflow from the formation.

GAS LIFT VERSUS PUMP-ASSISTED LIFT


One of the most basic decisions in selecting an artificial lift method is the choice between gas lift
and pump-assisted lift. If we are thinking about installing continuous gas lift, we first must consider
the physical limit gas-liquid ratio (GLR) and the optimal GLR.

PHYSICAL LIMIT GLR


The higher the rate of lift gas injection, the more we improve well performanceup to a point.
Consider the pressure traverse curves of (Figure 1). These curves were generated for a specified
set of flowing well conditions at a tubing depth of 4000 ft.

Figure 1

We can see from this figure that:


Adding gas to the liquid column lowers the hydrostatic pressure in the wellbore. In this case,
increasing the GLR from zero to 100 SCF/Bbl reduces the bottomhole pressure by nearly
600 psi.
Higher GLRs result in higher friction pressure losses, which offset the hydrostatic pressure
drop. For example, if we again increase the GLRthis time from 100 to 200 SCF/Bblthe
bottomhole pressure decreases by only 270 psi. As we further increase the GLR, the
corresponding bottomhole pressure decrease is smaller still.
Eventually, we reach a physical limit GLR where the increase in friction pressure becomes
approximately equal to the decrease in hydrostatic pressure. The net change in bottomhole
pressure becomes negligible. In (Figure 1), this occurs between 800 and 1000 SCF/Bbl. At
this point, injecting additional gas will not result in additional liquid production.

The bottomhole pressure corresponding to the physical limit GLR is likely to be much higher than
the bottomhole pressure we could attain by installing a downhole pump. This consideration alone
would tend to favor pump-assisted lift. But there are other factors to take into account. If a
formation is subject to drawdown-related production problems such as water coning or sand
production, for instance, then gas lift may be the preferred option.

OPTIMAL GLR
The choice between gas lift and pump-assisted lift ultimately comes down to economics. Back
when the costs of gas re-injection were insignificant compared with the benefits of increasing the
incremental oil production rate, the physical limit GLR was considered optimal for meeting gas lift
requirements.
Since that time, however, with production costs escalating and natural gas becoming valuable in
its own right, this is no longer a safe assumption. The actual optimum will be the GLR above which
the incremental cost of injecting additional gas exceeds the incremental revenue from increased
oil production. This may be lower than the physical limit value.

GENERAL DESIGN CONSIDERATIONS


Continuous gas lift design follows a systems analysis approach, in which pressures at various key
points are determined for the desired production rate and different GLR values. The sequences of
steps may vary, depending on which system parameters are known, and which are to be
determined. The two most fundamental design issues are

How much gas to inject?


At what depth(s) to inject it?

To address these issues, we must be able to determine how a well is likely to perform under
different operating conditions.

FORMATION DELIVERABILITY
A wells IPR defines the rate at which the formation can deliver fluid to the wellbore under a given
reservoir pressure drawdown. This relationship should be established for the current reservoir
pressure and, if necessary, for anticipated future reservoir pressures. We can then use the IPR to
determine what flowing bottomhole pressure is needed to maintain a desired production rate.
The flowing bottomhole pressure attainable from continuous gas lift may be constrained by the
formations tendency toward drawdown-sensitive problems such as sand production and water or
gas coning. In any case, it will be restricted to some minimum value by the physical limit GLR (or,
if it is lower, the optimal GLR) as described above. The limiting GLR, in turn, will be influenced by
the tubing head pressure required to deliver fluid from the wellhead to the separator, and on the
pressure losses that take place in the production tubing.
As the average reservoir pressure decreases over time, so will the gas injection pressure required
to produce at a given flow rate. This effect must be included in the gas lift system design.
Likewise, the system must accommodate increases or decreases in GLR or water cut. If we can
determine the magnitude of these changes, we must include them in our system design. If not, it
may be necessary to adjust the design or rely on subsequent wireline operations to modify valve
setting and placement.

TUBING HEAD PRESSURE REQUIREMENTS


The separator typically represents the downstream end of the production system. To determine the
flowing tubing pressure (FTP) needed for formation fluids and lift gas to flow to the separator, we
start with the known separator pressure and calculate the pressure losses that occur in the
gathering lines at the surface, between the separator and the wellhead.

TUBING DIAMETER
In most areas, safety and environmental regulations require that fluid be produced through the
tubing. In some fields with high-rate producing wells, however, annular production may be an
option.

(Figure 2) illustrates the difference in production performance for a specific well, based on the size
of the production tubing or tubing/casing annulus. The producing GLR is 400 SCF/Bbl, the flowing
tubing head pressure is fixed at 250 psi and the well's IPR has been calculated. We consider four
different production tubing installations:
1. Tubular flow through 2-inch tubing
2. Tubular flow through 2 1/2-inch tubing
3. Annular flow through 2 x 4 1/2- tubing-casing diameters
4. Annular flow through 2 1/2 x 5 1/2-inch tubing-casing diameters

Figure 2

For each curve, a higher flow rate results in a higher bottomhole pressure, and the performance
curves shift to the right with increases in the cross-sectional area available for flow. This means
that for a given bottomhole pressure, there is a significant flow rate increase with increasing tubing
diameter. The intersection of each tubing performance curve with the IPR curve specifies the
maximum production rate for the given GLR and surface tubing pressure. (If the desired
production rate is lower than this value, it is possible to use smaller production tubing with a
performance curve that intersects the IPR curve at a higher bottomhole pressure.) In this example,
the highest production rates would be attained through annular flow.

GAS INJECTION RATE

The gas injection rate frequired for continuous gas lift is estimated as follows:
(qgas)inj = 0.001x (GLR2 -GLR1)qliquid

(2)

where (qgas)inj = gas injection rate, MCF/D


GLR2 = producing gas-liquid ratio above the point of gas injection, SCF/Bbl
GLR1 = natural producing gas-liquid ratio (i.e., below the point of gas
injection), SCF/Bbl
qliquid = desired liquid production rate, B/D
For example, assume that a well's IPR and pressure traverse relationships indicate that it can
produce 850 B/D of fluid at a GLR of 1490 SCF/Bbl, and that the formation GLR is 452 SCF/Bbl.
The required gas injection rate would be
(qgas)inj = 0.001 x (1490 -452)850 = 882.3 MCF/D
A preliminary step in any gas lift system design is to inventory current and anticipated future gas
volumes. The injection gas will probably come from field production operations (usually from a
high-pressure separator). Additional makeup gas may be available from local gathering systems
or pipelines.

GAS INJECTION PRESSURE


The surface gas injection pressure that the system requires will depend on the gas lift design
parameters for individual wells, including the expected production rates, gas lift GLRs and the
depths of the operating valves. Keep in mind that increasing the producing GLR increases flowline
pressure losses and can affect separator performance.
Gas injection pressure requirements are based on the injection pressure required to open the
operating valve. This pressure is greater than the surface injection pressure by an amount equal to
the gas gradient times the injection depth. For a tubing installation under static conditions, we may
express the pressure gradient in the tubing-casing annulus as

(3)
where (Pinj)valve is the injection pressure at the operating valve,
(Pinj)surf is the minimum operating injection pressure at surface
D is the valve depth
We may obtain this gradient using the following equation, which is based on a mechanical energy
balance and the real gas law (Economides et. al, 1994):

(4)

where

= gas gravity (air = 1)


zavg = compressibility factor at average temperature and pressure
Tavg = average temperature, R = F +460

Because the compressibility factor depends on pressure and temperature, we must use an
iterative procedure to determine the unknown pressure. Where the casing pressure at the injection
point is known, we may obtain (Pinj) surf by trial-and-error as follows:
1. Assume values for (Pinj) surf and zavg at an arithmetic average pressure.
2. Use this value of zavg to calculate a new (Pinj) surf value from Equation 2.
3. Repeat steps 1 and 2 until the values of (Pinj) surf and zavg remain consistent within an
acceptable margin.

(This same procedure applies if (Pinj) valve is the unknown quantity and (Pinj) surf is known)
Gilbert (1954) introduced a simple approximation for Equation 3, in which he assumed = 0.7, zavg
= 0.9 and T = 600R, and then applied a a Taylor Series expansion:

(5)

An existing compressor, a high-pressure separator or gas from outside sources may provide
sufficient pressure for gas injection. If not, then the system design will have to be based on the
surface pressure that is available, or additional compression capabilities will have to be built into
the system.

LOCATION AND DESIGN OF THE OPERATING VALVE


In a tubing installation, lift gas from the surface enters the production tubing through the operating
valve. The operating valve is placed such that the gas injection pressure at the valve depth, minus
the pressure differential across the valve, is equal to the flowing production pressure at that depth.
Depending on the gas lift GLR, we can place the operating valve at any of several depths. It
requires less compression horsepower to inject gas at a high pressure and low rate than it does to
inject it at a low pressure and high rate. Thus, a gas lift design operating at a low GLR requires a
higher operating pressure and a lower compression horsepower. By selecting the lowest operating
GLR, we can minimize the horsepower and gas volume requirements for a given production rate.
If the available gas volumes or injection pressures are limited, that limitation should be considered
in the design. If, on the other hand, we are constrained only by the physical limit GLR, then our
design recommendations will be governed by the above guidelines and by the economics of gas
injection.
The design of the operating valve depends on the type of operation and any anticipated fluctuation
in flow rate, producing GLR and water cut. To provide a constant GLR when there are varying
production rates, the operating valve should exhibit the gas throughput characteristics of a
throttling valve. The capability to adjust the valve through wireline operations allows modification of
the design to accommodate changing conditions.
Tubing diameters should be specified to give the lowest operating GLR for the target production
rate. This will minimize horsepower requirements, surface operating pressure, and the volume of
injected gas.

SYSTEM DESIGN EXAMPLE


Consider a well for which we want to establish a production rate of 600 B/D, and for which we
have the following information:

Depth: 5600 ft to midpoint of perforations (MPP)


Average reservoir pressure = 1750 psi at MPP
Target production rate = 600 B/D (32 API oil; 20% water cut)
Formation (current) GLR = 180 SCF/Bbl
Pwf required for target rate = 914 psi, based on IPR analysis
Average reservoir temperature = 147 F
Average surface temperature = 65 F
Tubing size: 2 3/8 inch OD
Required tubing head pressure: 100 psi
Assume a pressure differential of 100 psi across the operating valve

DESIGN OBJECTIVE
We want to incorporate this well into an existing gas lift system. The injection line pressure is 800
psi, and the injection pressure available at the wellhead will be 750 psi. Our goal at this stage of
the design is to determine
(1) the depth of the operating valve and
(2) the required gas injection volume.
The operating valve should be placed to minimize the gas injection volumes and compression
horsepower requirements. This means that we will want to inject at the maximum available
pressure of 750 psi.

DESIGN PROCEDURE
The basic design procedure is adapted from Brown (1980, vol. 2a).

Pressure Traverse
Using an appropriate multi-phase vertical flow correlationin this case, a spreadsheet-generated
model based on a modified Hagedorn-Brown correlation (Economides et. al, 1994) and starting
at bottom with the 914 psi required to produce 600 B/D at a GLR of 180 SCF/Bbl, we establish
the pressure traverse relationship for the current well conditions ( Figure 3: Pressure traverse
curve for example well, 600 B/D at 180 SCF/Bbl GLR).

Figure 3

In this well, the pressure in the tubing goes to zero before it reaches the surface. In other words,
the well cannot produce at the desired rate of 600 B/D under natural flowing conditions. This well
is a candidate for gas lift.

Gas Gradient and Operating Valve Depth


Using Equation 4 and the iterative procedure described above, we can generate the gas injection
gradient in the casing-tubing annulus (Figure 4: Gas injection gradient for example well, surface
injection pressure of 750 psi).

Figure 4

The depth at which the gas injection pressure is equal to the tubing pressure is referred to as the
point of balance. It is represented graphically by the intersection of the pressure traverse curve
and the gas injection gradient; in this example, the point of balance is at 5372 ft.
To account for the 100 psi pressure differential across the operating valve, the valve needs to be
set above the point of balance. The setting depth is represented graphically by the intersection of:

the tubing pressure traverse curve for GLR = 180 SCF/Bbl, and
a line parallel to the gas gradient curve and separated by 100 psi.

The depth at which these intersect is 4940 ft (Figure 5: Determination of gas injection point for
example well). This is the lowest depth at which we can place the operating valve.

Figure 5

Alternatively, we could determine the operating valve depth by generating pressure traverse
curves for a set of gas lift GLRs ranging from just above the current 180 SCF/Bbl to the physical
limit GLR. For each gas lift GLR, the point of injection would be the depth at which its pressure
traverse intersected the curve for the current GLRin other words, where the tubing pressure
above the injection point equals the tubing pressure below the injection point
In this case, because our design objective is to minimize the gas injection volume and compressor
horsepower requirements at a given injection pressure, we are using the gas injection gradient to
define the location of the operating valve.

Gas Injection Rate


To determine the gas injection rate, we must find a GLR for which the pressure traverse fits these
two endpoints:

The flowing bottomhole pressure at the injection point of 4940 ft


The flowing tubing pressure of 100 psi.

Using a spreadsheet-based, trial-and-error calculation, we determine that the optimal GLR is 500
SCF/Bbl (Figure 6: Determination of optimal gas lift GLR for example well) . Thus, the GLR will be
500 SCF/Bbl above the injection point, and 180 SCF/Bbl below the injection point.

Figure 6

From Equation 2, we can now determine the gas injection rate as


(qgas)inj = 0.001(500 -180)600 = 192 MCF/D
Again, this gas lift installation could have been designed for a higher GLR and a lower injection
pressure, in which case the operating valve would have been set at a shallower depth. Selecting a
lower GLR, however, fits in with our stated design objective for an established gas injection
pressure.

Production Rate Variations


The gas lift valve, located at the injection point, must have a large enough opening to handle the
required gas volume at a casing-tubing pressure differential of 100 psi. But what would happen if
the flow rate dropped from 600 B/D to 300 B/D? Assuming critical flow conditions, a fixed-size
orifice valve would allow the same volume of injection gas to flow for each production rate.
Therefore, the GLR at the 300 B/D rate would be too high, resulting in inefficient fluid production.
Alternatively, if the production rate was to increase from 600 B/D to 900 B/Dfor example,
following a successful well stimulation treatmentthe fixed-size orifice valve would be too small to
deliver enough gas to the fluid column.
When variable production rates are possible, a gas lift valve must be able to respond to these
variations while delivering sufficient gas to provide a constant GLR.
Where the inflow performance is expected to change over time, a throttling valve may be used to
allow a proportional response to flow rate. If this type of valve is used, it should be dynamically
performance-tested before installation, to ensure that that the desired gas flow rates at the various

casing and tubing pressures will actually occur. The flow testing of valves is often carried out at the
manufacturer's shop under simulated operating conditions. The throttling valve, however, does
have certain disadvantages:

Because throttling valves have limited gas throughput capacity, high-volume wells sometimes
require an additional valve to reach the desired lift depth.

Pressure adjustments are more difficult than with certain other types of gas lift valves
Since more information must be considered, throttling valve designs are more technically
detailed and thus require a higher degree of accuracy.

UNLOADING IN CONTINUOUS GAS LIFT WELLS


Unloading or kicking off is the process of removing a static column of liquid from a wellbore so that
the well can flow. A well must be unloaded when either of two conditions occurs:

Liquid has accumulated in the well and the fluid level has reached a static point below the
surface.

The well is full of kill fluid following completion or workover operations, and is ready to be put
on production.

In either case, the well will not produce until it is unloaded. This unloading process must be part of
the gas lift design.
In completion and workover operations, we typically unload wells by swabbing, injecting nitrogen
or circulating a fluid of lower density than the kill fluid. In continuous gas lift, we assume that the
gas injection rate and pressure used to produce the well is available for unloading (in some
facilities, an additional kick-off pressure may be available); all that will be needed are additional
valves in the tubing, spaced at appropriate depths for unloading fluids.
Because of the hydrostatic pressure that the kill fluid exerts in the tubing-casing annulus, the
operating valve is open, and the fluid levels in the casing and tubing are equal prior to unloading.
Normal practice for unloading gas lift wells is to set the top unloading valve just above the static
fluid level. Because most wells that produce on continuous gas lift have relatively high formation
pressures, their static fluid levels tend to be at or near the surface. For design purposes, therefore,
we may assume that the fluid column reaches to surface.
If the density of the wellbore fluid is known, then that value should be used to determine the
unloading pressure gradient. If the fluid density is unknown, assume that its gradient is slightly
higher than a fresh water gradient of 0.433 psi/ft - 0.45 psi/ft is a good value for design purposes.

GENERAL PRINCIPLES
Unloading calculations are based on providing sufficient pressure at progressively lower valve
depths to U-tube liquids to the surface, through each valve in succession from top to bottom.
To illustrate how the process works, consider a gas lift well with three valves in the tubing string.
From top to bottom, these are

Valve 1 (unloading valve)


Valve 2 (unloading valve)
Valve 3 (operating valve)

Before unloading, the static fluid level is at the surface, and all valves are open. Unloading takes
place according to the following sequence:

1. Gas is injected into the annulus at the unloading pressure. The gas displaces the liquid in the
annulus through the open valves and into the tubing. The liquid is u-tubed to the surface.
2. As liquid is displaced, the fluid level in the annulus drops until Valve 1 is uncovered.
3. Gas begins to flow into the tubing through Valve 1, reducing the liquid density in the tubing
above the point of injection and decreasing the tubing pressure.
4. Because Valves 2 and 3 are still open, the annular fluid level continues to drop as gas is
injected through Valve 1 and the pressure in the tubing decreases.
5. The tubing pressure drops to a stable value as the annular fluid level reaches a corresponding
stable depth. If the unloading string is properly designed, this stable fluid level should be just
below Valve 2.
6. Valve 1 should close just as gas begins to flow through Valve 2. This makes Valve 2 the sole
point of gas injection.
7. As gas is injected through Valve 2, the density of the liquid in the tubing above Valve 2 is
reduced. The tubing pressure decreases, and the liquid level in the annulus falls because
Valve 3 is still open.
8. The tubing pressure drops to a stable value as the annular fluid level reaches a corresponding
stable depth. If the unloading string is properly designed, this stable fluid level should be just
deep enough to uncover Valve 3.
9. At this point, Valve 2 closes and Valve 3, the operating valve, becomes the sole point of gas
injection.

UNLOADING DESIGN PROCEDURES


The basic objective in designing an unloading string is to progressively transfer the point of gas
injection downhole, from the top to the bottom unloading valves, and finally to the operating valve.
Each valve closes in turn as the next deepest valve is uncovered, and all of the valves above the
operating valve remain closed during normal production.
Beyond these general considerations, design procedures vary according to the valve type and to
the completeness and reliability of the design input data.
Takcs (2005) describes several representative valve spacing methods developed for various
types of valves. These include procedures for:

Injection pressure-operated (IPO) valves:


o Variable pressure drop per valve (Winkler and Smith, 1962)
o Constant pressure drop per valve (API RP 11V6, 1999)
o Constant surface opening pressure (U.S. Industries, 1959)
Balanced valves
Production pressure-operated valves
Throttling valves

The design example that follows is based on using IPO valves with constant surface opening
pressures. In this procedure, the top-to-bottom injection transfer is accomplished by increasing the
production (i.e., tubing) pressures (U.S. Industries, 1959).

SYSTEM DESIGN EXAMPLE


Table 1 summarizes the design parameters that have been established for a well to be placed on
continuous gas lift.
Table 1: System Design Example Continuous Gas Lift
Average reservoir pressure

1750 psi at midpoint of perforations (5600 ft)

Pwf required for target rate

914 psi, based on IPR analysis

Target production rate

600 B/D
(32 API oil, water cut = 20 percent )

Average reservoir temperature

147 F

Average surface temperature

65 F

Formation GLR

180 SCF/Bbl

Tubing:

2 3/8 inch

Flowing tubing pressure (FTP)

100 psi

Gas lift GLR

500 SCF/Bbl

Gas injection rate

192 MCF/D

Depth of operating valve

4940 ft

Tubing pressure at operating valve


depth

744 psi

Kill fluid

Lease crude oil, kill gradient = 0.38 psi/ft; assume


fluid to surface. Formation overbalance = 378 psi

As a starting point, we refer to (Figure 6). This shows the gas injection and design gradients for
this well, along with the pressure traverse curve above and below the point of injection.

Figure 6

For design safety, and to minimize valve interference, it is typical to select a surface design
pressure that is the greater of these two values:

FTP + 200 psi, or


FTP + 0.2 x (Injection pressure)

In this example, the larger of these values is 100 + (0.2 x 750) = 250 psi. Using this surface design
pressure and the tubing pressure at the operating valve depth, we can generate a tubing design
pressure gradient as shown in (Figure 7).

Figure 7

The depth to the top unloading valve (D1) may be calculated using the relationship

(1)
where (Pinj) = gas injection pressure, psi
FTP = flowing tubing pressure, psi

kill = kill fluid gradient, psi/ft


g = average gas gradient, psi/ftin this example, the average gas gradient is estimated
at 0.0183 psi/ft.
In this case,

This result may also be obtained graphically by starting from the FTP and generating a line parallel
to the kill fluid gradient. The depth at which this line intersects the gas injection pressure gradient
curve corresponds to D1. From the tubing design gradient line, we determine that the tubing design
pressure at this depth is 430 psi. (Figure 8: Unloading designlocation of top unloading valve.)

Figure 8

The tubing design pressure at D1 represents the starting point for locating the second and
subsequent valves. Again, we use the pressure gradient of the kill fluid to determine the required
gas design pressure. (Figure 9) illustrates the process graphically.

Figure 9

For this example, five unloading valves are needed to unload the well, along with the operating
valve to provide continuous gas lift operations:
Valve No.

Depth

1797 ft

2772 ft

3533 ft

4124 ft

4584 ft

6 (Operating valve)

4940 ft

There are a number of design options for locating the unloading valves, especially when
considering the available selection of valves and the ability to specify the valve opening and
closing pressures. Often, for example, a lower valve will be opened before an upper valve is
closed in order to provide a smoother unloading operation.

CONTINUOUS GAS LIFT: SUMMARY DESIGN


PROCEDURE
Below is a summary design procedure for a continuous gas lift installation. Depending on the
design program in use, certain of these steps will be carried out automatically once the input
parameters are entered. Also, the sequence of steps may vary depending on which design
parameters (e.g., surface injection pressure, maximum gas volumes, etc.), if any, are already
defined.

1. Collect all available input data; Table 1 provides a starting point for listing input
parameters.
Table 1: Input Data for Continuous Gas Lift Design
Well number, field, location, zone

Casing data

Average reservoir pressure

Tubing and production equipment data

Average bottomhole, surface


temperatures

Directional profile

Anticipated pressure decline rate

Perforation density, diameter

Back pressure or separator


pressure

Sand control equipment

Gas-oil ratio, water cut

Current gas lift configuration (for existing


installations)

Gas specific gravity, oil API gravity

Available gas injection rate, surface injection


pressure
(for existing surface facilities)

Water specific gravity

Kill fluid density

Type of well completion

Preferred correlations for vertical flow

Depths of perforations

Preferred correlations for surface flow

2.
3. Establish the wells Inflow Performance Relationship for the current reservoir pressure
and, if desired, anticipated future reservoir pressures.
4. Carry out a production system analysis to define pressure traverse curves for various
combinations of producing GLR and production tubing assemblies. For each
combination, generate the pressure traverse by evaluating the lift performance in the
tubing string and the flowing tubing head pressure needed to move fluid from the
wellhead to the production separator. For analysis purposes,
the range of tubing diameters is based on the diameter of the production
casing and/or the type of completion (if annular flow is permitted,
consider it as a design option)

5.
6.

7.
8.

the lower GLR limit is the formation GLR, while the upper GLR limit is the
physical limit GLR above which additional gas injection will not
increase the wells production rate.
Combine these pressure traverse calculations with the IPR to establish the optimal
production rate.
For the selected tubing or annular flow system, determine the depths at which the
pressure traverse curves for various values of gas lift GLR intersect the pressure traverse
curve that corresponds to the wells natural flow. These intersections define several
possible locations for the operating valve.
Select the depth for the operating valve, keeping in mind that the lowest GLR requires the
highest operating pressure, the lowest injected gas volume and lowest compression
horsepower. This would normally be the option selected.
Determine the injection pressure as a function of depth. At the operating valve, this
pressure will be equal to the required production pressure at that depth plus the pressure
corresponding to the valve spread. With this pressure as a starting point, obtain the gas
gradient in a static gas column using the equation below or a suitable approximation or
correlation:

where:
(Pinj)valve is the injection pressure at the operating valve, psia
(Pinj)surf is the minimum operating injection pressure at surface, psia
g
= gas gravity (air = 1)
h
= operating valve depth, feet
zavg = compressibility factor at average temperature and pressure
Tavg = average temperature, R
Note: Where the lift system is being designed from the bottom up, and (Pinj)valve has been determined, this
equation is used to determine the minimum required gas injection pressure at the surface. For an existing
gas lift system, where (Pinj)surf is already established, this relationship will be used to determine (Pinj)valve j for
various depths.

9. Locate the unloading valves using the appropriate procedure for given valve types and
operating conditions.
10. Review the manufacturers specifications to select the unloading and operating valves
that will satisfy the system design for the life of the well. The values are set to meet the
required operating conditions.
11. Before completing the system design, be sure there is sufficient gas volume and pressure
to meet operating conditions and, finally, be sure that the design can be modified as
inflow performance characteristics change.
A gas lift system designer must realize that the design is not static. The inflow performance,
producing GLR and water cut of the well will probably change as production from the well
continues. Also, a system is often designed for an entire field rather than a single well. The future
performance of individual wells that collectively make up the system must be considered when
developing the optimal field design.

INTERMITTENT GAS LIFT


Intermittent gas lift is similar to continuous gas lift in that it employs some of the same equipment
and procedures, and it improves well performance through the injection of pressurized gas from
the surface into the production tubing. The way in which it works, however, is very different.

Continuous gas lift is based on introducing a constant stream of gas into the producing fluid
column in order to reduce its density.

Intermittent lift, on the other hand, is based on periodically injecting slugs of compressed gas
below the liquid that accumulates in the production tubing. These gas slugs do not change
the liquid density; instead, they physically displace or push the liquid to surface. (Figure 1:
Intermittent gas lift, closed installation).

Figure 1

Intermittent lift involves three steps, or periods (Figure 2: Intermittent gas lift cycle):

Figure 2

1. Inflow or Accumulation: Fluid from the formation flows into the wellbore and collects in the
tubing above the gas lift valve. This period continues until a sufficient volume of liquid has
accumulated, based on the wells inflow performance and the system design. At this point,
the hydrostatic pressure of the accumulated fluid causes the standing valve to close.
2. Lift: The operating valve opens, and a high-pressure slug of gas flows into the tubing,
displacing the accumulated liquid up the tubing and to the surface. The lift period ends when
the last of the liquid has moved into the flowline.
3. Pressure Reduction or Afterflow: As the liquid exits the well, the hydrostatic head in the
tubing is reduced, and the pressure of the gas slug rapidly dissipates. When the pressure
drops below a certain level, the operating valve closes, the standing valve opens and
another inflow period begins.

Control of the intermittent lift cycle is based on regulating the frequency and duration of the lift
period, along with the amount of gas injected during the lift period. The optimal system design is
one that combines maximum liquid recovery with an economical volume of injected gas.

ESTIMATING PRODUCTION CAPABILITY


The production capability of an intermittent gas lift system depends on three factors:

Starting load
Lift efficiency
Number of cycles per day

STARTING LOAD
The starting load reflects the pressure at the operating valve just as the valve opens. It is the
pressure exerted on the operating valve by the column of fluid in the tubing above it.
A typical design might employ a Starting Load Factor of 50 to 75 percentthat is, fluid will be
allowed to build up in the tubing until the tubing pressure is 50 to 75 percent of the available
casing pressure. For a normal design situation, a smaller load factor would be used; when there is
a high surface tubing pressure or a high gas delivery rate into the tubing, the Starting Load Factor
would be closer to 75 percent. In either case, the excess casing pressure provides the slug
velocity.
For a given starting load, we may calculate the volume of liquid in the tubing as follows:
p = pt - pts
h = p/Gs
Be = (h)Ftb

(1)

where: p = pressure imposed by the fluid in the tubing above the valve.
pt = bottomhole tubing pressure
pts = surface tubing pressure
h = height of rise of fluid in the tubing. Ignoring the gas column in the tubing, it is equal to p
divided by the pressure gradient of the produced liquid, G s
Ftb = tubing volume factor, volume/unit of length (e.g., Bbl/ft)
Be = fluid influx volume/cycle (the volume of fluid in the tubing available for lift during each cycle
e.g., Bbl/cycle).

LIFT EFFICIENCY
The next consideration is whether the entire volume of fluid inflow into the tubing is lifted during a
cycle and, if not, what level of efficiency exists.
As a slug of fluid moves up the tubing during lift, some of the fluid adheres to the tubing walls and
some becomes entrained as droplets in the gas phase. This lost fluid is referred to as holdup.
Field tests have shown that a holdup of five percent to seven percent of the starting load per 1000
feet of lift will exist when the starting load is within 65 to 75 percent of available casing pressure
(Winkler and Smith, 1962). These conditions exist when the slug velocity is optimum and holdup is
low.
For an assumed loss of 5 percent per 1000 feet of vertical lift, the efficiency of lift, E, will be equal
to:

(2)
where:

E = efficiency of lift, percent

Dv = depth to the gas lift valve, feet.


If, for example, the gas lift valve is at a depth of 4000 feet, the efficiency of lift will be:

With this information, we may calculate Bt , the liquid produced per cycle:

(3)
Under these conditions, 80% of the starting load should be produced. With this information, we
can calculate the volume of fluid produced per cycle.

NUMBER OF CYCLES
Knowing the volume of fluid produced per cycle, the next step is to calculate the number of lift
cycles that are possible per day as follows:

(4)
where Nc = number of cycles per day
tc = minimum time per cycle, minutes per 1000 ft of depth
Dv= depth of valve, ft
Nc depends on the depth of lift and the length of time required for pressure reduction and inflow
periods. The cycle time is usually adjusted in the field under actual operating conditions, but initial
estimates may be made.
For example, given a particular range of operating conditions, a reasonable first assumption for
the minimum time per cycle might be on the order of 3.0 minutes per 1000 feet of lift (Winkler and
Smith, 1962). For an operating valve located at a depth of 4000 feet and a minimum cycle time
factor of 3 minutes per 1000 feet of depth, then, the maximum number of cycles per day is:

The daily production rate is the product of Nc and Bt. This estimate provides a good starting point
for determining daily production. However, when the intermittent system is installed in the field, it
may turn out that the ideal number of cycles per day, as determined through field tests, is less than
this maximum.
To estimate the maximum daily production rate q, use the following equation:

SAMPLE PROBLEM
Estimate the production capability of a well under the following conditions:

Depth of the operating valve = 8000 feet


Tubing size = 2 3/8-inches OD
Surface tubing pressure = 100 psi

Surface operating gas pressure = 800 psi


Gas gravity = 0.65
Oil gradient = 0.40 psi/ft

Calculate the maximum daily production rate of the well using intermittent gas lift operations.
1. First, calculate the liquid inflow per cycle.
2. Determine the gas pressure in the annulus opposite the operating valve, with a known surface
pressure of 800 psi, and an estimated gas gradient of 0.02125 psi/ft. The gas pressure at the
operating valve is
pc = 800 + 0.02125 (8000) = 970 psi.
Using a 65 percent load factor, find p

p
pt
p

= pt - pts
= 0.65 pc = 0.65 (970) = 630 psi
= 630 - 100 = 530 psi.

With this pressure, the produced fluid should rise in the tubing to a total height of:

This is equivalent to a fluid inflow volume of:


Ftb

= 0.0038 bbl/ft (from tubing tables or calculations)

Be

= hFtb

=(1325) (0.0038) = 5.03 bbls.

3. Next, calculate the lift efficiency.


For a depth of 8000 feet and an assumed 5% holdup per 1000 ft, the efficiency, E, is

With a lift efficiency of 60%, the fluid production per cycle is:
Bt = 0.60 (5.03) = 3.02 bbls/cycle.
4. Next, calculate the maximum number of cycles possible per day. For a gas lift operating valve
located at a depth of 8000 feet, it is equal to:

5. Complete the calculations by combining the production per cycle and the maximum number of
cycles per day to find a maximum production rate of:
q = NcBt = (60) (3.02) = 180 B/D.
If the inflow performance calculations for this well show that the well will sustain this rate, it is very
likely a good design. However, some degree of field adjustment will probably be required.

VALVE SELECTION

The primary requirement of an operating valve used in intermittent lift is to be able to handle a
large volume of gas over a short time period. This is an ideal application for the pilot-operated
valve. Its large port allows a large volume of gas to pass once the valve is opened. A properly
designed dome-charged valve, or a fluid-operated valve, could also be used. However, because
the port of a fluid-operated valve is small, a series of operating valves, opening in succession,
would be required to move the slug up the tubing.Required Gas VolumesTwo-phase slug flow is a
complex phenomenon, and as such, it is difficult to calculate the exact gas volumes required
during an intermittent cycle. For estimation purposes, we may assume that the required volume
equals the volume of gas left in the tubing at the moment that the slug reaches the surface, and
that the pressure of the gas in the tubing is equal to the average of the two values of tubing
pressure when the valve opens and closes.The basic gas volume required per cycle is equal to:

where:
pt = pressure at the operating valve;
pvc = the pressure just as the valve closes;
Vt = the volume of tubing not occupied by fluid, and;
pa
= atmospheric pressure (used to convert gas volume in the tubing to standard
conditions), in this example, 14.73 psia.
This estimate is approximate, and does not include the effects of temperature and compressibility.)
For example, continuing with the Sample Problem from above, and given the following
information for this 180 B/D well,

Valve opening pressure is 970 psi


Valve closing pressure is 725 psi
Tubing length is 8000 feet; capacity = 0.0217 cubic feet per foot.
Fluid fills 1325 feet of the tubing.
The tubing gas volume is V t = (8000 - 1325) (0.0217) = 144.8 cu. ft.

We can estimate the gas volume per cycle as follows (converting to standard cubic feet of gas in
the tubing at the average pressure):

This agrees with a common rule-of-thumb whereby the injection gas requirement can be
approximated as 200-400 SCF/Bbl per 1000 ft of lift for a well on conventional intermittent lift
(Winkler and Smith, 1962):

This gas must be conserved and reused in subsequent cycles to create an efficient production
operation.

SURFACE CONTROL OF INJECTED GAS

The gas lift valve controls the flow of gas from the casing to the tubing. Surface controllers, which
may be actuated on a time-cycle or by a choke, complement the subsurface valve control function.
With time-cycle control, a clock drives a pilot that opens and closes a diaphragm-actuated valve
on the gas supply line. The pilot can be adjusted to open and close for specific time periods.
The choke control uses the inflow performance of the well and the operating spread
characteristics of the gas lift valve to control the cycle. The surface equipment consists of an
adjustable choke or flow control valve on the gas supply line. The choke is adjusted to admit gas
continuously into the annulus so that its pressure builds at a steady rate.
When the pressure reaches a specified level, the gas lift valve opens and the slug is displaced.
When the choke is set to admit gas at a rate compatible with the wells inflow capacity, an efficient
cycle frequency is established. A very important feature of choke control is that it eliminates the
cyclical injection surges from the compressor and effectively isolates the cyclic surges to the
casing annulus. The compressor operates more evenly and the gas circulated to the well can thus
be measured more accurately.
The difference between the opening and closing pressures of the gas lift valve is called its spread.
This feature of choke control allows the casing annulus to store the volume of gas needed for
each intermittent lift cycle. In other words, the gas lift valve spread makes the casing annulus act
as a storage chamber.

UNLOADING INTERMITTENT GAS LIFT WELLS


Intermittent gas lift design is similar to continuous gas lift design in that liquids are U-tubed to the
surface from one valve to the next. Therefore, the procedures for locating the unloading valves in
a continuous lift system generally apply to intermittent lift as well. The one significant difference for
intermittent flow is that we must define a pressure gradient in the tubing string for times when kill
fluids are unloaded as slugs. Under these conditions, the pressure gradient in the tubing is caused
primarily by frictional losses that are a function of the velocity of flow and tubing size. Empirical
correlations indicate that these design gradients range from 0.02 to 0.35 psi/ft. We may use the
appropriate gradient from these correlations to generate the design gradient and find the optimal
locations for the unloading valves.

PRACTICAL ASPECTS OF WELL UNLOADING AND


OPERATION
The guidelines in this section are adapted primarily from API RP 11 V5 (1999), Recommended
Practice for Operations, Maintenance and Troubleshooting of Gas Lift Installations.

INITIAL UNLOADING
The first step in bringing a well on production after gas lift valves have been installed is to unload
the fluids from the wellbore and obtain a stabilized production rate. Normally, a well placed on
continuous gas lift is unloaded continuously, and a well placed on intermittent gas lift is unloaded
intermittently. Primary considerations in unloading include avoiding excessive pressures that could
damage the valves, and using clean, filtered workover fluids to avoid plugging or abrasion of the
valves.
Prior to unloading, a two-pen pressure recorder should be installed at the surface to monitor both
the gas injection pressure and the production (tubing) pressure. These pressures should be
measured as close to the wellhead as practical. In any case, the gas injection pressure should be
measured downstream of the injection choke, and the production pressure should be measured
upstream of any flowline choke that is present. The wellhead pressure should be bled down to the
pressure of the downstream separator, and the flowline choke, if present, should be either fully
open or removed.

CONTINUOUS GAS LIFT WELLS

With continuous gas lift, the unloading process begins when gas is injected slowly into the
annulus, probably through a choke located at the surface. Pressure is incrementally raised by
approximately 50 psi every eight to ten minutes until it reaches about 400 psi, and 100 psi every
eight to ten minutes thereafter. The kill fluid is displaced through the standing valve, up the tubing
and to the surface into a disposal tank, until gas starts coming around the first valve or oil appears
in the produced fluid. A steady stream of fluid will be then unloaded. If these fluids are directed into
a separator, it is important to keep the backpressure on the well as low as possible. As gas is
continuously injected into the annulus, a gradual increase in casing pressure will be required to
keep fluids flowing from the tubing string.
Valve 1, the uppermost valve, is the first valve to be uncovered; gas first enters the tubing string at
this point. This is noted at the surface by an immediate increase in the velocity of the fluid stream
coming out of the tubing. A mixture of gas and liquid is soon produced at the surface, and the
casing pressure levels off at the surface operating pressure of Valve 1. As gas continues to enter
the annulus, the liquid column in the annulus is lowered until Valve 2 is uncovered. As soon as this
valve is uncovered, gas flows through it and enters the tubing. Casing pressure then drops to the
surface operating pressure of this valve. At about the same time, pressure in the annulus opposite
Valve 1 should have been reduced to a level low enough to cause that valve to close.
Unloading continues from valve to valve until the deepest operating valve is uncovered. At this
point, the bottomhole pressure has been reduced to a level that allows the formation fluid to move
into the tubing, and the volume of gas injected through the operating valve is sufficient to lift the
production under design conditions.

INTERMITTENT GAS LIFT WELLS


With intermittent gas lift, fluid is unloaded at the surface in the form of piston-like slugs. The
unloading process is the same as that for continuous flow until the uppermost valve (Valve 1) is
uncovered. At that point, the well is placed on intermittent control for unloading. This is
accomplished with a choke or a time-cycle controller at the surface so that the well is alternately
produced and shut-in. During this period, the fluid in the annular space will continue to be U-tubed
into the tubing and will be produced as slugs. A good cycle for unloading is obtained with 2 to 4
minutes of gas injection every 20 to 30 minutes. This allows ample time for stabilization to take
place between slugs.
When the well is unloaded down to the operating valve, the choke size or cycle time should be
adjusted to suit the wells production characteristics. Thus the unloading operation may start with a
high number of cycles per day and then, in response to the wells production behavior, the number
of cycles will be adjusted downward as fewer cycles will be needed to maintain optimum
production rates. If the fluid production rate begins to fall off, then the number of daily cycles is too
low for optimal production. With this information, it is possible to make further refinements to the
process by reducing the duration of gas injection during each cycle. The objective is to maximize
production and minimize the gas volume required. A very useful monitoring procedure involves
simultaneously recording the shapes of the tubing and casing pressures curves. Adjustments are
made on the basis of the shapes of these two curves.

SYSTEM ADJUSTMENTS
Once a well is unloaded, the next step is to optimize its production rate and gas usage. This will
require some adjustment of its operating parameters. For detailed procedures, refer to API RP
11V5 (1999).
In a continuous gas lift installation, adjustments are generally made using an adjustable choke to
control the rate of gas injection (a positive choke could also be used, but this would require
interrupting gas injection to change the choke size). To prevent freezing, the gas system may be
equipped with a dehydrator, gas heater or heat exchanger, or methanol may be injected upstream
of the choke. To adjust the gas injection rate, the choke is initially set at a diameter that is larger
than required for the design rate. The diameter is reduced incrementally until the production rate
begins to drop, and then readjusted to establish the optimal production rate.

Similar types of adjustments are made for intermittent gas lift installations that employ time cycle
control: the controller is initially set for a duration that will exceed the design gas injection
requirements, and then the number of cycles per day is reduced until the well can no longer
produce at its desired rate. The controller is then reset in steps until the optimal production and
gas injection rates are established. For intermittent wells operating on choke control, the choke is
initially sized for the design production rate, and then adjusted in the same type of trial-and-error
manner.

GAS LIFT SYSTEM MONITORING


Successful gas lift performance depends largely on the efforts of field personnel. A gas lift
installation requires close supervision during the unloading process, and when injection gas is
adjusted and regulated.
A common practice is to analyze the system only when a problem arises. A better approach is to
analyze each well while it is operating satisfactorily to determine if the installation has been
properly designed. This provides a baseline measure of performance for reference in the event of
trouble, and helps to indicate needed design changes. It is important to analyze this baseline
information before planning well servicing or workover operationsotherwise, the operator will not
know what changes are needed.
Diagnostic tools for monitoring and troubleshooting gas lift wells include:

Two-pen pressure recorder charts and calibrated pressure gauges installed at the well
Acoustical and production logging surveys
Fluid level determinations using wireline

API RP 11V5 (1999) describes these tools and their applications in detail.

REMOTE MONITORING
Gas lift wells are a common area of application for remote monitoring and control techniques.
These systems can measure the performance of a single well or an entire field using sensors and
data transmitting devices that alert field personnel to changes in well performance. By comparing
performance parameters over time, the operator can analyze well stability, allocate lift gas
injection, and optimize the operation of the entire field. This capability leads to improved efficiency,
better field operations management, and increased profitability.
A significant feature of monitoring systems is their ability to remotely control gas injection and
change well settings using two-way control devices. Continuous monitoring and comparison of
parameters such as injection pressure, wellhead pressure, and flow rate lets the operator identify
potential problems and take preventive and corrective action from a central location. In many
cases, the operator is notified automatically when sensors detect significant changes to key
parameters.
Primary components of remote monitoring systems include:

Downhole Pressure & Temperature Sensors


Sensor Data Process System
Well Controller
Remote Terminal Unit (RTU)

The downhole pressure and temperature sensors communicate via a system controller to adjust
gas injection through the sensor data process system. This allows the operator to control the well
or the field based on changing surface or downhole conditions.

A remote terminal unit (RTU) can transmit data continuously or store well performance data for
later transmission and analysis. The RTU is a two-way system, thus allowing the operator to
communicate back to the well.
The monitoring and communication equipment is powered by solar cells, which are backed up by
a battery system to ensure a constant power supply.
Through monitoring of gas lift injection and production systems, field efficiency can be improved
and future gas lift valve design, valve placement and unloading programs can be designed on the
basis of actual field operating experience.

GAS LIFT SURFACE FACILITIES


(Figure 1) shows the main elements of a surface gas lift system, beginning with production at the
wellhead and ending with the injection of gas into the casing annulus or tubing.

Figure 1

Starting at the wellhead, produced fluids travel first to the separator. The separator gas is usually
re-used as lift gas. If more gas is produced than is needed for gas lift, the excess gas is either sold
or re-injected into the formation.
Moving downstream, there is a point where outside supply may be added to the system if the gas
from the separator is not sufficient to meet the demand. Both the separator gas and the outside
makeup gas flow through a scrubber, where impurities are removed.
The gas next moves to the compressor, where gas pressure is raised to desired levels. The
compressor must provide the appropriate discharge pressure and volume needed at both average
and peak rates. Some of the gas reaching the compressor is normally used as fuel.
Downstream of the compressor, gas is metered and various controls are introduced before the gas
is injected into the annulus. In the case of continuous injection, the control is normally a choke in
series with a pressure regulator. For intermittent gas injection, a time-cycle controller or choke is
the most common form of control.
It is clear that the surface system consists of a number of individual components, each of which
must be designed to provide the quantities and peak demands of gas for the gas lift system at the
desired injection pressures. It is also easy to see that the ideal gas lift system, especially with
respect to the compressor operation, is one that has a constant suction pressure and constant

discharge pressure on the compressor. This is easy to achieve in continuous flow operations,
because of the continuous supply of gas available from the separator and because of the need to
continuously inject gas. Intermittent systems are more complicated with intermittent injection and
production - the duration of which may vary for each well. Control is more difficult with time-cycle
control than for choke control because with the choke, the annulus serves as a storage chamber
between lift cycles.

DATA COLLECTION
The first step in designing surface facilities for a gas lift installation is to collect the following data:

Number and location of wells requiring gas lift


Gas lift valve design for each well
Whether continuous or intermittent injection will be used
Gas volumes needed (along with estimates of peak demand)
Availability of gas supply from the separator or external supply
Location of sales gas lines
Pressure required at the point of injection into the well
Pressure of the separator or supply gas
Sizing of the compressor
Auxiliary control and metering system required for the surface system.

CALCULATING COMPRESSOR HORSEPOWER


The compressor is a major component of a gas lift system. Compressors are available in many
different sizes and horsepower ratings to handle different gas lift operating conditions.
An approximation for determining a compressors brake horsepower (bhp), is given by the
equation

(1)

where

n
= number of stages
Q
= gas throughput capacity, MMCFD (106SCF/D)
(Pdischarge/Psuction) = overall absolute compression ratio
1.05
=correction factor for pressure drop and gas cooling between stages

(For a single-stage compressor, the correction factor reduces to 1.0)

The quantity
, which represents the absolute compression ratio per stage, should
not be greater than 4.
Example:
Find the horsepower needed to move 2500 MCFD (2.5 MMCFD) through a 2-stage
compressor with 50 psi suction pressure and 200 psi discharge pressure.
First, find the absolute compression ratio per stage:

Continue with Equation 1 as follows:


bhp = 1.05x 23 x 2 x 2 x 2.5 = 241.5 hp => use a 250 hp compressor
To solve the problem using a single stage compressor, the solution would be:
bhp = 1.00 x 23 x 1 x 2.5 .4 = 230 hp.

DESIGN SAFETY FACTORS


For the surface gas lift system to have enough capacity, it is customary to estimate a mainline
pressure that is approximately 100 psi higher than is called for in the design. This additional
pressure will accommodate unexpected line losses. In addition, the compressor delivery volume is
usually increased by 10% to account for volume losses and the fuel needed for compression.

SUMMARY OF DESIGN PROCEDURES


We may summarize the procedure for designing a surface system for gas lift installation as
follows:
1. Begin by laying out the entire surface system, including the wells, gathering lines, stock tanks,
separators, and other items of equipment that materially affect gas lift operations.
2. Specify the wells that will use either continuous or intermittent gas lift. Also consider the time
during which each well will use gas lift.
3. Design the gas lift system for each well, specifying the pressures, the volume, the cycles, and
the expected life of the gas lift operation for that well. This is a key element of the design
because it provides the pressures, volumes, and cycles to which the system must respond
over time.
4. Make production estimates, including the gas volumes and pressures that will be available
from the separator. These volumes and pressures serve as input for the compressor
calculations.
5. Specify the gas sales and makeup volumes needed, along with their availability.
6. Design the balance of the surface system, including the gathering lines and the control
system.
7. Design the system compressor. A reasonably accurate measure of the required horsepower
can be calculated; however, it is advisable to discuss these estimates with the manufacturer
to ensure that the final design will meet the needs of the gas lift system.
8. Finally, remember to include a volume safety factor of 10% and a pressure safety factor of 100
psi.

This procedure can be used in a preliminary design of a gas lift system. This preliminary design
should be reviewed with the representatives of a gas lift equipment manufacturer. The optimized
final design will be one that satisfies all the requirements of the system without being overdesigned.

Reciprocating Rod Pump Systems


ROD PUMPING OVERVIEW

(Figure 1) shows the different parts of a sucker rod pumping system, including (from the top
down) its five major components: the prime mover, which provides power to the system; the gear
reducer, which reduces the speed of the prime mover to a suitable pumping speed; the pumping
unit, which translates the rotating motion of the gear reducer and prime mover into a reciprocating
motion; the sucker rod string, which is located inside the production tubing, and which transmits
the reciprocating motion of the pumping unit to the subsurface pump; and the subsurface pump.

Figure 1

(Figure 2) is a cross-sectional representation of a subsurface pump at two different stages of the


pump cycle. Note the location of the standing valve at the bottom of the tubing, and the traveling
valve at the bottom of the sucker rods. Note also the changing position of the plunger.

Figure 2

On the left-hand side of the figure , the plunger is approaching the bottom of its downstroke.
The traveling valve is open, and so the standing valve is closed because it is carrying the
weight of the fluid above it. At this point in the cycle, the fluid above the standing valve is
moving upward through the open traveling valve.

On the right-hand side of the figure , the plunger has reached the bottom of the stroke and is
just beginning to move upward. The plunger starts to lift the weight of the fluid above it, and
the traveling valve closes. As the plunger continues to move upward, the volume in the
working barrelbetween the standing valve and traveling valveincreases, while the
pressure in the working barrel decreases. As soon as this pressure becomes less than the
flowing bottomhole pressure, the standing valve opens and formation fluids flow upward.
During each upward movement of the plunger, wellbore fluids are lifted a distance equal to
one full stroke length.

When the plunger reaches the top of its stroke, its movement is reversedthe traveling valve
opens, the standing valve closes and the cycle repeats its reciprocating movement of the rods and
the opening and closing of the two valves. With each stroke, fluid is moved up the tubing toward
the surface.
If the produced fluid contains free gas, there are two points to note regarding the pump cycle:
1. The valves will not necessarily open and close at the exact top and bottom of the stroke. The
point in the upstroke at which the standing valve opens will depend on the spacing (the
volume that exists at the bottom of the stroke, between the traveling and standing valves),
and on the amount of free gas present in this volume. On the downstroke, the traveling valve
remains closed until the pressure below the plunger is greater than the pressure above it.
The traveling valve then opens and allows fluid to pass through it into the tubing. The exact
point in the downstroke at which the traveling valve opens depends on the free gas volume
in the fluid below the valve.
2. The greater the volume of free gas, the greater the proportion of the stroke that is taken up in
gas expansion and compression, without any true pumping action taking place.
For wells producing a reasonable volume of gas, a gas anchor is normally installed on the tubing
below the pump. This device allows the separated gas to be produced up the annulus before it
would otherwise enter the pump.

ADVANTAGES AND DISADVANTAGES OF ROD PUMPING


SYSTEMS
Rod pumping systems can be used to reduce bottomhole pressures to very low levels, and offer
great flexibility for low-to-medium production rates. They are relatively simple with respect to
design, operation and maintenance, and can be adapted to a wide range of operating conditions.
They account for the large majority of artificial lift wells, and are one of the most well-known and
generally understood systems in the field. Surface and downhole equipment can easily be
refurbished, and tends to have high salvage values.
While they have a very wide range of applications, rod pumping systems are mainly limited to
onshore locations due to the weight and space requirements of surface pumping units. Solids
production, corrosion and paraffin tendencies, high gas-liquid ratios, wellbore deviation and depth
limitations due to sucker rod capabilities have all been seen as problem areas for this lift method,
although some of these problems have been alleviated due to improvements in sucker rod
metallurgy, the development of long-stroke pumping units and other technical advances.

SURFACE EQUIPMENT
The primary surface equipment components of a rod pumping system are the prime mover, the
gear reducer and the pumping unit (Figure 1: Courtesy Lufkin Industries, Inc).

Figure 1

PRIME MOVER
The prime mover, which may be either an internal-combustion engine or an electric motor,
provides power to the pumping unit. The choice of prime mover for a particular well depends on
the field conditions and type of power available.

INTERNAL-COMBUSTION ENGINES
Internal-combustion engines are most the most commonly used prime movers. They are classified
as either slow-speed or high-speed engines.

Slow-speed engines have one or two cylinders and usually operate at speeds of up to
750 RPM.
High-speed engines have four or six cylinders; they can operate at speeds ranging from
750 to 2000 RPM, but are usually run at less than 1,400 RPM. Generally, high-speed
engines have lower torque than slow-speed engines of comparable horsepower.

Engines can be further classified as two-cycle or four-cycle engines.

Most two-cycle engines are one-or two-cylinder, slow-speed engines. Two-cycle engines
operate on natural gas, liquid petroleum (LP) gas, or diesel fuel.

The four-cycle engine is manufactured in either slow- or high-speed versions.


A slow-speed, four-cycle engine has a single horizontal cylinder and uses a large flywheel to
provide a constant speed to the pumping unit. Some slow-speed, single-cylinder engines run
on diesel or fuel oil, although these fuels are not often used if natural gas is readily available.

Four-cycle, high-speed engines used as prime movers can be fueled by natural gas, LP gas
or gasoline.

Five factors determine which type of internal combustion engine to use:

Available Fuel - Natural gas is often supplied to the engine from the wellhead or from a field

separator. This gas may require processing to remove oil, water and acid gas components
such as H2S or CO2. Separator gas is considered the best fuel source, although LPG, diesel
fuel and light-gravity crude oil have also been used.
Equipment Life and Cost - Slow-speed engines, although they may cost more initially, tend
to have lower maintenance costs and longer operating lives than high-speed engines.
Engine Safety Controls - Because these engines run unattended, they must have reliable
safety controls. Most safety controls stop the engine in cases of high water temperature, low
oil pressure, overspeed, or pumping unit vibration (in case of sucker rod breaks). The
engines are usually stopped by grounding the magneto and shutting off the fuel.
Horsepower - Power requirements are dictated by factors such as the size of the pumping
unit, the depth of the well and the fluid gravity. API publication API 7B/11C describes
procedures for the testing and rating of engines.
Installation - Slow-speed engines must be set on heavy concrete foundations. Multi-cylinder
or vertical engines can be set on much lighter foundations.

ELECTRIC MOTORS
Induction electric motors are also used to drive pumping units. Horsepower ratings range from 1 to
200 HP, but most motors on pumping units operate at 10 to 75 HP. These are most often threephase motors.
In fields where three-phase power is not available, single-phase AC motors may be used to meet
power requirements of up to 10 HP on shallow, low-volume stripper wells. However, these
motors cost more and operate less efficiently than three-phase motors having similar ratings.
Direct current (DC) motors may also be used to drive pumping units, but they are not often chosen
due to their higher initial and maintenance costs, their inability to use electricity from utilities and
the fact that DC voltage cannot be changed by transformers, making transmission and distribution
difficult.

Motor Selection
Oversized electric motors are often installed to prevent under-powering of pumping equipment.
This may reduce motor failures, but it also increases capital costs and power consumption.
When a new pumping unit is installed on a well, motor sizing is based on preliminary estimates of
the depth and size of the pump, speed of the pumping unit, and fluid properties. Actual motor
requirements are determined by variables such as fluid level, viscosity, well deviation, and quality
of electric power, as well as friction in the pump, stuffing box, and pumping unit. As a result, it is
difficult to accurately size a motor for a new installation on the first attempt. Sometimes multiplerated motors are considered for pumping unit drivers because they can be modified to operate at
different horsepower settings.

Motor Controls
Motors used for pumping units have two types of control devices: one that stops, starts, and
controls the motor, and the other that protects the motor. Common control devices include:

The Hand-Off-Auto Switch, which is used to shut off the motor or operate it in automatic or
manual mode.

o
o
o

In the hand position, the motor runs continuously, but does not override
protection controls.
In the off position, the motor stops and cannot start, but power is not shut off.
In the automatic position, a programmer or time clock controls the motor.
Pump-off controls usually function only when the switch is in auto position.

The Line Disconnect Switch, which cuts all electrical power to the motor and is used when
maintenance is performed on the pumping unit.

The Sequence-Restart Timer , which delays the simultaneous restarting of several motors. If a
power failure occurs in the field, a severe voltage drop can result as many motors connected
to the same power source try to simultaneously restart. This condition could potentially keep
the motors from restarting or cause the unintended operation of other field safety control
devices.
The Programmer, which controls the running times of pumping units. If the pumping unit
operates continuously, the pump may not fill completely on the upstroke. This pumped-off
condition will cause fluid pounding as the pump strikes the top of the fluid column on the
downstroke. Such pounding causes shock loading of the sucker rods and pumping unit.
Programmers optimize the pumping unit run time and reduce damage to the pumping
system.

Devices that are designed to protect oilfield motors include:

Motor fuses, which are located between the motor control and the electric power and limit
damage to the motor from electrical or mechanical problems.

Air circuit breakers, which are used instead of fuses to protect the electrical distribution
system in case the motor develops an electrical problem.

Lightning arrestors, which protect the system against storm damage caused by lightning.
Undervoltage relays, phase loss relays, thermal overload relays, and motor winding
temperature sensors, which are used to stop the motor before any permanent damage is
done.
Pumping hit vibration switches, which shut down the pumping unit motor if excessive vibration
is detected. Such vibrations tend to occur when sucker rods break.

Electrical Distribution System


The distribution system that supplies power to oil field electrical systems is made up of 3 main
components: the primary and secondary electrical systems, and the grounding system.
The primary system supplies electricity to the field at voltages ranging from 4000 to 15000 V. Highvoltage primary systems are preferred when electrical power travels a long distance. To protect
these systems, it is necessary to install lightning arrestors and static lines that dissipate high static
charges associated with electrical storms.
High voltages from the primary system are transformed to lower voltages (generally less than 600
V) through a secondary system. The secondary system includes the distribution transformers that
convert the primary system voltage to the motor operating voltage and all cables, switches,
controls, and other devices that operate at the motor voltage.
Electrical equipment at the well site must be properly grounded for personnel safety and to ensure
that the electrical devices perform satisfactorily.

If utility-furnished power is not available, field-based generators can provide electrical power to
operate pumping unit motors . Distribution equipment for the generator system would be the same
as for utility power. The voltage requirements will depend on the size of the producing field .

In a field with fewer than five wells located close to the generator, the generator voltage may
be the same as the motor rated voltage.

In a field with 5 to 50 wells, the generated voltage should be higher than the motor rated
voltage. At each motor, a transformer would be used. This system would be considered a
moderately sized system, with a generator having a distribution of 2,300 or 4,160 V.
In a very large field with more than 50 wells, generated voltage of 2,300 or 4,160 V would be
stepped up to 7,200 or 13,800 V for distribution. The higher voltage allows the use of smaller
conductors and reduces line voltage drop. At each wellsite, a transformer will drop the
distribution voltage to motor rated voltage.

API Specification 11L6, Electric Motor Prime Mover for Beam Pumping Unit Service (1993,
supplemented 1996) covers motor specifications and requirements for motors of up to 200 hp.

GEAR REDUCER
The gear reducer is used to convert the high speed and low torque generated by the prime mover
into the low speed and high torque required by the pumping unit. The Lufkin gear reducer shown
in Figure 2 is of the double reduction type (Figure 2: Double reduction gear reducer. Courtesy
Lufkin Industries, Inc) . A high-speed gear (the smallest in diameter) is mounted on a shaft, which
is connected to a sheave/belt assembly that is driven by the prime mover. Speed reduction occurs
between this gear and a larger gear mounted on an intermediate shaft, and between the
intermediate gear and a still larger gear mounted on the crankshaft that actually drives the
pumping unit. The gears are continually lubricated by an oil reservoir contained within the
assembly. When in place at the well, the gear train is mounted in an enclosed box.

Figure 2

PUMPING UNIT
The pumping unit changes the rotational motion of the prime mover to a reciprocating vertical
motion. The unit is driven by the crankshaft on the gear reducer, and is connected to a polished

rod and a sucker rod string, which drives the subsurface pump. Most pumping units employ a
counterbalance (usually adjustable weights or pressurized air), which opposes the weight of the
sucker rod string.
Pumping units are available in a variety of sizes and configurations. They are classified according
to their methods of counterbalance and they ways in which their major components are arranged.

CLASS I LEVER SYSTEMS


Class I lever systems include the conventional crank-balanced, beam-balanced, and certain
special geometry units. The fulcrum (Samson post bearing) of this type unit is located at the
middle of the walking beam, between the well load and the actuating force ( Figure 3: Class I lever
system: conventional pumping unit. Courtesy Lufkin Industries, Inc).

Figure 3

Conventional Crank-Balanced Units


As the cranks on a conventional unit rotate, the pitman side members cause the walking beam to
pivot on a center bearing, moving the polished rod. Adjustable counterweights are located on the
cranks (Figure 4: Conventional crank balance pumping unit geometry ).

Figure 4

This is the most common pumping unit type, because of its relative simplicity of operation, low
maintenance requirements and adaptability to a wide range of field applications.

Beam-Balanced Units
On beam-balanced units, the counterweights are positioned at the end of the walking beam
(Figure 5: Churchill beam-balanced pumping unit. Courtesy Lufkin Industries, Inc). This type of
unit is susceptible to damage at high pumping speeds, and so its use is limited to smaller sizes
and low pumping speeds. Nonetheless, its simple design and dependability make it an attractive
option for shallow wells.

Figure 5

Crank Balanced Units with Special Geometry


On some crank balanced units, the gear reducer is moved from directly under the equalizer to a
position away from the centerline of the well. This difference from the conventional geometry
changes the torque factors and the time intervals on the upstrokes and downstrokes. Units of this
type usually have an out-of-phase counterbalance system and require a specific rotation direction.

CLASS III LEVER SYSTEMS


Class III lever systems include air-balanced and the Mark II crank counterbalanced units
(Figure 6: Class III lever system: air-balanced unit. Courtesy of Lufkin Industries, Inc. The walking
beam hinge point is at the rear of the unit and the actuating force is located between the pivot
point and the well.

Figure 6

Air-Balanced Units
Air-balanced units are similar to crank-balanced units in that the rotation of the crank causes the
walking beam to pivot and move the polished rod (Figure 7: Air-balanced pumping unit
geometry) .

Figure 7

A piston and air cylinder partially counterbalances the well load by compressing air in the
cylinders. An auxiliary air compressor, controlled by a pressure switch, maintains the system air
pressure (Figure 8: Air balance piston/cylinder assembly. Courtesy of Lufkin Industries, Inc .).

Figure 8

The piston/air cylinder assembly allows for more accurate control of the counterbalance than the
use of counterweights, and reduces the weight of the unit. This results in lower installation and
transportation costs . Air balanced units work to particular advantage on wells that require larger
unit sizes or longer pump strokes, where space and weight requirements might preclude using a
crank-balanced unit.

Mark II Units
On Mark-II units, the cross yoke bearing is located very close to the horsehead ( Figure 9: Mark II
pumping unit. Courtesy of Lufkin Industries, Inc.). The cranks, which rotate in only one direction,
have an angular offset to produce an out-of-phase condition between the torque exerted by the
well load and the torque exerted by the counterbalance weights. These features reduce problems
associated with torque peaks that are more common with conventional crank balanced units.

Figure 9

LONG-STROKE PUMPING UNITS


For a given depth, pump size, production rate and rod string, a longer pump stroke and slower
pumping speed tend to result in reduced power requirements and lower rod stresses than an
equivalent system operating with a shorter pump stroke and a higher pumping speed (Tait and
Hamilton, 1984). This observation is particularly worth noting in deep, high-volume wells that might
normally call for an electrical submersible or hydraulic pump. Long-stroke pumping units (e.g.,
Weatherford International's Rotaflex unit) provide another option. Where conventional pumping
units are available with maximum stroke lengths of 240 inches, these units have stroke lengths of
up to 366 inches, operating at rates ranging from 1 to 5.2 strokes per minute (Lea, Winkler and
Snyder, 2006). Under the right conditions, these units can be used to optimize lift efficiency,
increase rod life and minimize power costs.

PUMPING UNIT SELECTION


The choice of pumping unit will be based on well parameters, operating conditions and, of course,
cost and availability. A conventional crank-balanced unit might be chosen because field personnel
are familiar with it, while the relatively small size, low weight and low inertial and shaking forces of
an air-balanced unit would make it a good choice for an offshore site or other confined surface
location.

PUMPING UNIT API DESIGNATIONS


Manufacturers design pumping units in standard sizes, and their catalogs feature pumping units
that can handle many combinations of torque, polished rod load and stroke length. Pumping units
are designated using a 10-character, alphanumeric code established by the API (Figure 10: API
pumping unit designations). An example of a pumping unit designation would be C-140D-117-64.

Figure 10

The first character in the API code designates the pumping unit type.
1.
2.
3.
4.
5.
6.
7.
8.

C - Conventional
A - Air-balanced
B - Beam-balanced
M - MARK II
RM - Reverse Mark
LP - Low Profile
CM - Portable/Trailer Mount
LC - Power Lift

The next four characters designate the peak torque rating, in thousands of inch-pounds, and
the type of gear reducer. In most cases, a double reduction gear reducer is used, indicated
by the letter D.
The next three characters define the peak polished rod load rating, in hundreds of pounds.
The last two characters are the stroke length, in inches.

Thus, the pumping unit C-160D-117-64 is a conventional unit with a peak torque of 160000 in-lb, a
double reduction gear reducer, a peak polished rod load of 11700 lb and a maximum stroke length
of 64 inches.
Table 1(after Lufkin, 2004) lists standard sizes and API designations for conventional crankbalanced pumping units:
Table 1
Standard Sizes for Conventional Crank-Balanced Pumping Units (after Lufkin, 2004)
C-1824D-305240

C-640D-365168

C-320D-256144

C-160D-173100

C-80D-119-64

C-1824D -365216

C-640D-305168

C-320D-256120

C-160D-17386

C-80D-135-54

C-1824D -365192

C-640D-365144

C-320D-213120

C-160D-20074

C-80D-119-54

C-1280D-305240

C-640D-305144

C-320D-305100

C-160D-17374

C-80D-133-48

C-1280D-427-

C-640D-256-

C-320D-256-

C-160D-143-

C-80D-119-48

192

144

100

100

C-1280D-365192

C-640D-365120

C-320D-24686

C-160D-17364

C-912D-305240

C-640D-305120

C-320D-21386

C-160D-14364

C-912D-365192

C-456D-305168

C-320D-24674

C-114D-11986

C-912D-305192

C-456D-305144

C-228D-213120

C-114D-14374

C-912D-365168

C-456D-265144

C-228D-213100

C-114D-17364

C-912D-305168

C-456D-365120

C-228D-173100

C-114D-14364

C-912D-427144

C-456D-305120

C-228D-24686

C-114D-17354

C-912D-365144

C-456D-256120

C-228D-21386

C-114D-13354

C-456D-213120

C-228D-20074

C-456D-256100

C-228D-17374

A manufacturers selection of pumping units may not exactly fit a particular application. In that
case, it is better to install a pumping unit that is slightly over-designed, rather than one with lesser
capabilities.

SUBSURFACE EQUIPMENT
The subsurface components of a rod pumping system include the sucker rod string and the
subsurface pump.

SUCKER RODS
The subsurface pump is connected to the pumping unit on the surface by a string of solid sucker
rods (Figure 1). These rods come in either 25-foot or 30-foot lengths, and API-standard diameters
of 1/2, 5/8, 3/4, 7/8, 1 and 1-1/8 inch.

Figure 1

Pony rods are shorter-length rods used to bring the rod string to the exact length needed in a well.
Sucker rods are joined together by threaded connectors or couplings, which are usually about 4
inches long (Figure 2: Courtesy of Tenaris Oilfield Services, 2003 -- http://www.tenaris.com)).
Some rods are manufactured with pin-and-box couplings, while others are made with pin
couplings at both ends and then joined together using rod couplings.

Figure 2

API Specification 11B (Twenty-sixth edition, 1998) provides dimensional and material standards
for sucker rods and couplings.

STEEL SUCKER RODS


One-piece steel sucker rods are manufactured from hot-rolled steel. The pin ends of the rod are
forged, the rod is heat-treated, the pins are machined, and a finish designed to reduce corrosion is
applied to the rod surface.
Three-piece rods are manufactured by attaching threaded metal pin and box connectors to the
threaded ends of a machined rod. These sucker rods are sometimes used in shallow wells.
The choice of what size and API grade of steel sucker rods to use in a well is determined by rod
stress and well conditions. Rod stress is determined by the production rate, tubing and pump sizes
and the characteristics of the pumping unit.
API RP 11L (1988, Reaffirmed January 2000) sets forth design calculation recommendations for
sucker rod pumping systems. A companion volume, API Bulletin 11L3, (1970, Reaffirmed
September 1999) provides tables of computer-calculated values for selecting sucker rod systems.
These tables are developed for various rod string, pump stroke, pump size and pumping speed
combinations.
The presence of H2S, CO2, salt water or other corrosive agents effectively reduce the maximum
allowable stress on a rod string, and require adjustments to rod stress calculations.

FIBERGLASS SUCKER RODS


Fiberglass sucker rods are actually made of fiberglass embedded in a plastic matrix. Steel end
fittings with standard API pins are attached using an adhesive. These rods are joined with
standard couplings to form rod strings.
The primary reasons for using fiberglass rods is that they lighter and more corrosion-resistant than
steel rods. This makes them an attractive option in deeper wells and corrosive environments.
There are, however, several disadvantages to using fiberglass rods in place of steel rods. They
are more susceptible to fatigue failures resulting from the repeated loading and unloading during
the pumping cycle, and are more likely to fail under compressive forces. And they may lose their
strength through exposure to high well temperatures or hot-oiling treatments. The resins used
during manufacture of the rods are the elements that determine the rods resistance to chemicals
and temperature; different resins have different physical properties and resistance to chemicals.
API standards for the manufacture and performance of fiberglass sucker rods are detailed in API
Specification 11B (1998), which covers dimensional requirements for fiberglass sucker rods,
couplings and subcouplings..

CONTINUOUS SUCKER RODS


Unlike conventional sucker rod strings, continuous sucker rods (e.g., COROD, provided by
Weatherford International Ltd) require couplings only at the top and bottom of the rod string regardless of well depth. This continuous length of steel reduces the weight of the rod string and
provides a more uniform contact with the production tubinga particular advantage in deviated
wells.

ROD FAILURES
Steel sucker rods usually break because of failures in the rod body, most often caused by
corrosion resulting from exposure to H2S, CO2, salt water, or O2. Corrosion inhibitors have been
developed to help reduce these types of rod failures.
Body breaks in steel rods commonly occur near the upset. A bending moment placed on the rod in
the area of the upset can weaken the rod. Sucker rods may also fail at the pin if too much torque
is applied when the rod string is being made up. Extra care when making up or breaking down and
handling sucker rods will reduce these types of failures.
Fiberglass rods usually fail at the end-connector joint where the metal end connector may
separate from the rod, although breaks or separations of the rod body also occur. Fiberglass rods
are more easily damaged than steel rods; surface scratches and ultraviolet light may weaken
them.
API Spec.11BR (1989, reaffirmed October, 1993), covers recommendations on the storage,
transportation, running and pulling of sucker rods.

SUBSURFACE PUMPS
The main components of a subsurface pump are:

The working barrel, connected to the tubing.


The plunger, connected to the sucker rods.
The traveling valve, which is part of the plunger assembly.
The standing valve, which is located at the bottom of the working barrel.

At the bottom of the pump, connected to the working barrel, there is usually a perforated gas
anchor, which allows formation fluids to separate before entering the pump. It also directs much of
the free gas into the casing-tubing annulus and improves pump efficiency.
(Figure 3) shows the two principal categories of subsurface pumps: the tubing pump on the left,
and the rod or insert pump ).

Figure 3

The main difference between these two pump types is in the installation of the working barrel.

The working barrel of a tubing pump is an integral part of the tubing string. This is
advantageous in that it provides for the strongest possible pump construction, and it allows
the maximum plunger diameter to be only slightly less than the tubing diameter, thus
maximizing the volume of fluid that can be pumped. The disadvantage of the tubing pump is
that the entire tubing string must be pulled in order to service the working barrel and other
pump hardware.
The working barrel of a rod pump is run on sucker rods rather than on tubing. Thus, the pump
can be serviced by pulling only the rods. However, its plunger diameter has to be smaller
than that of a comparable tubing pump, and thus it cannot move as large a fluid volume.

TUBING PUMPS
Tubing pumps are classified according to the type of working barrel, the standing valve
arrangement and the type of plunger used.

Working Barrel
The working barrel may be either a one-piece, tube-type barrel, or a liner contained within an
outside jacket (Figure 4: Working barrels).

Figure 4

The one-piece working barrel, which may be heavy-walled or thin-walled, is made of cold-drawn,
seamless steel, cast iron, or corrosion resistant alloys. The working barrel is polished on the inside
to allow easier plunger movement.
The widely used common working barrel is a heavy-wall version of the one-piece barrel.
The liner-type working barrel consists of a liner surrounded by an outer steel jacket. The liner may
consist of a single hardened steel liner or several short sectional liners held in place by clamping
collars (Figure 3). The liner-type working barrel has greater precision and lower repair cost, but
higher capital cost than the common working barrel.

Standing Valve
The second criterion for classifying a tubing pump is whether its standing valve is fixed or
removable.

(Figure 5) shows a tubing pump with a fixed standing valve. In part (a), the standing valve is
located at the bottom of the tubing string; in part (b), the plunger is pulling out and the standing
valve is still located at the shoe of the tubing. To retrieve the standing valve, the tubing string must
be pulled, as shown in part (c).

Figure 5

(Figure 6) shows a tubing pump with a removable standing valve. The standing valve is located at
the shoe of the tubing when the pump is operating. To remove the valve (b), the plunger is lowered
onto a special fitting on the standing valve, which is then brought to the surface along with the
plunger and sucker rods. The barrel and tubing are left in place in the subsurface.

Figure 6

Plunger Type
A third criterion for classifying pumps is the makeup and composition of the plunger. There are two
types of plungers: the cup-type (or soft pack) plunger, and the metal plunger.

The cup-type plunger is the oldest form of seal used in pumping units. Early cups were usually
made of leather or rubber-impregnated canvas; however, synthetic materials are now used
for flexible plungers. On the upstroke, pressure exerted by the fluid column forces the cup to
expand and form a seal between the lip of the cup and the wall of the barrel. On the
downstroke,, as pressure is equalized on both sides of the cup, it collapses inward to allow
the plunger to fall freely., These types of plungers are not generally used below 5,000 ft.
Metal plungers (Figure 7) are made of cast iron or steel and have either a smooth or a
grooved sealing surface. Grooved plungers offer an advantage when the well produces

sand. Sand particles can be trapped in the groves, preventing them from abrading the
plunger. The metal-to-metal seal for these plungers depends upon an extremely close
clearance. This type of plunger usually wears better than the cup-type and is used in deeper
wells.

Figure 7

Note that it is possible to have both cup-type and ring-type plungers in a single pump.

ROD PUMPS
The two main types of rod pumps are the stationary pump (Figure 8), which has a moving plunger
and a stationary working barrel, and the traveling barrel pump (Figure 9), which has a moving
working barrel and a stationary plunger.

Figure 8

Figure 9

Another less common type of rod pump is the three-tube pump. This pump uses an inner plunger
and an outside barrel that telescope down in a concentric manner around a standing barrel, thus
forming a long fluid seal between the barrels. Because this pump does not use mechanical seals,
it performs well with abrasive fluids in sandy wells.
The casing pump (Figure 10) is a special type of rod pump in which a packer, placed at either the
top or bottom of the working barrel, provides the fluid pack-off between the working barrel and the
casing. No tubing is used. Casing pumps are generally used in large volume, shallow pumping
applications.

Figure 10

API PUMP CLASSIFICATIONS


The API classification for subsurface pumps, set out in API Specification 11AX (2001), is based on
the specification of three elements of the pump:
1. Whether it is a rod-type or tubing-type pump
2. Whether it has a stationary or traveling working barrel and the type of barrel and plunger
installed
3. Whether it has a top or bottom seating assembly (for rod pumps with stationary barrels), or
anchor (an anchor is a mechanism that keeps one section of the pump stationary so that the
pump can operate properly. It is also referred to as a hold-down ).

The API classification uses a three-letter code. Each letter refers to one of the three elements
above. The full range of API pump classifications is shown in ( Figure 11) (API pump
classifications. After API Spec. 11AX) :

Figure 11

Pumps (a) and (b) in this figure are examples of stationary barrel, top hold-down pumps. Top
hold-down pumps are recommended for use in sandy wells, because sand particles cannot
settle over the seating nipple. This makes the pump easier to remove. These pumps are
also good for wells with low fluid levels, because the standing valve is submerged deeper
than it would be with a bottom hold-down pump. Pump (a) has the API classification RHA.

R means that it is a rod-type pump,

o
o

H signifies that it has a stationary, heavy-wall barrel, and


A is used to indicate a top anchor.

The top anchor refers to the fact that the standing valve is held in place from above. If the
stationary, heavy-wall barrel is changed to a stationary, liner barrel, the designation for the
pump changes to RLA. In this case, only the middle letter changes in our pump
classification.

Pump (b) is similar to the type of RHA pump. Again, it is a rod-type pump, but it has a

stationary, thin-wall barrel and a top anchor. Its designation is RWA. If a soft-packed plunger
is installed, then the designation changes to RSA.
A rod-type pump with a stationary, heavy-wall barrel and a bottom anchor is an RHB pump
(pump (c) in the figure). Note that for a bottom anchor pump, the standing valve is held from
below and a (B) designation, rather than an (A) designation is used. If the barrel is changed
to a stationary liner barrel, the designation changes to RLB.
An RWB pump (pump (d)) has the same configuration, but with stationary, thin-wall barrel. If a
soft-pack plunger is used in the pump, the designation is RSB./li>
A traveling, heavy-wall barrel with a bottom anchor on a rod pump (pump (e)), is designated
RHT. If the traveling barrel is a liner barrel, the designation is RLT.
Pump (f) is the same type of pump, but with a thin-wall traveling barrel; its designation is
RWT. If a soft plunger is added, the designation is RST.
If, instead of a rod-type pump, a tubing pump with a heavy-wall barrel is used, then the
designation is TH (pump (g) ). (If a liner barrel replaces the heavy-wall barrel, the
designation is TL.
A tubing pump with a heavy-wall barrel and a soft-pack plunger is given the designation TP
(pump (h) in Figure 12). Note that a metal plunger is used in situations where a specific
designation is not given.

NOTE: There is no particular need to try and memorize each of the sixteen designations just
covered. It is sufficient to know that the API classification exists and how to use it. It should be
noted that, only three or four of these designations are generally used.

Standard Designations for a Complete Pump


Building on the general classifications above, the API has developed standard designations for a
complete pump (API Specification 11AX, 2001). The designation consists of a 12 character,
alphanumeric code (XX-XXX-X-X-X-X-X-X-X) as shown in Table 1(after API 11AR, 2000):
TABLE 1: API Designation for Subsurface Pumps
(after API 11AR, 2000)
Characters
1,2

3, 4, 5

XX

Designation/Description
9

10

11

12
Tubing size:
15 - 1.900 inch O.D.
20 - 2.375 inch O.D.
25 - 2.875 inch O.D.
30 - 3.500 inch O.D.

XXX

Pump bore (basic):

TABLE 1: API Designation for Subsurface Pumps


(after API 11AR, 2000)
Characters
1,2

3, 4, 5

Designation/Description
9

10

11

12
106 - 1.0625 inch
125 - 1.25 inch
150 - 1.50 inch
175 - 1.75 inch
178 - 1.78125 inch
200 - 2.00 inch
225 - 2.25 inch
250 - 2.50 inch
275 - 2.75 inch

Type pump:
R - Rod
T - Tubing
X

Type of barrel:
(For metal plunger pumps)
H - Heavy wall
L - Liner barrel
W - Thin wall
(For soft-packed plunger
pumps)
S - Thin wall
P - Heavy wall
X

Location of seating assembly:


A - Top
B - Bottom
T - Bottom, traveling barrel
X

Type of seating assembly:


C - Cup type
M - Mechanical type
X

Barrel length, feet


X

Nominal plunger length, feet


X

Total length of extensions,


whole feet

To illustrate the use of the API designation, we may use the information from Table 1 to describe
an API pump designation for a rod-type, stationary pump with a heavy-wall barrel, a metal plunger,
and a top anchor using a mechanical-type seating assembly, to be set in 2.375-inch tubing, with a
pump-bore size of 1.5 inches. The plunger-length is 4 feet and there are no extensions to be
added.

The first two characters refer to a tubing size code. The 1.9-inch O.D. tubing, for example, is
code number 15; a 3 1/2-inch O.D. tubing is code number 30. In this example, with 2-3/8inch O.D. tubing installed in the pump, so the code number 20 are the first characters of
the pump designation.
The next three characters refer to the code for the pump bore size. There are nine pump bore
sizes and each has its own three-number code referring to sizes ranging from 1 1/16 inches
through 2 3/4 inches in diameter. In this example, a bore size of 1 1/2 inches, the next three
characters will be the number 150.
The next four characters refer to a description of the pump.

o
o
o

The first character specifies whether the pump is a rod or tubing pump. The
example pump is a rod type pump, designated by the letter R.
The second character refers to the type of barrel and specifies whether it is a
heavy-wall, a thin-wall or a liner barrel. The example pump has a stationary,
heavy-wall barrel, designated by the letter H.
The third character refers to the location of the seating assembly, whether top
or bottom, and if on the bottom, whether it is a stationary or traveling barrel.
The example pump has a top anchor, and this is designated by using the
letter A.
The fourth character in this series refers to the type of seating assembly.
There are only two choices: a cup-type assembly, indicated by the letter C,
or a mechanical seating assembly, indicated by an M. The example pump
has a mechanical assembly so the letter M is placed in the fourth position.
A rod-type pump with a stationary, heavy-wall barrel, a metal plunger, and a
top anchor with a mechanical-type seating assembly will have the designation
RHAM.

The next three characters refer to pump length. The first character refers to the barrel length
in feet. For this example, it is 8 ft. The second character refers to the nominal plunger length,
again in feet. A rule of thumb states that to limit fluid slippage, pump plungers should be:

o
o
o

3 feet long in wells less than 3,000 feet deep


3 feet plus 1 foot in length for each 1,000 feet between 3,000 and 6,000 feet
of depth
6 feet long for wells 6,000 feet and deeper
One exception to this rule is that shorter plungers should be used for pumping
viscous oils. For thisexample, assume a plunger length of 3 feet.

The final character is the total length of extensions to the barrel, again expressed in feet.
Extensions of 6 inches to 4 feet can be added to both ends of a heavy-wall barrel. These
extensions are used primarily to prevent scale buildup within the barrel. For the example
pump, assume an extension length of 2 feet.

Therefore, for this example, the final API pump designation would be 20-150-RHAM-8-3-2.

In order to purchase a pump, the only additional information needed would be the type of material
that the liner or barrel and plunger are to be made of, the plunger clearance, the valve material,
and the length of each extension to be installed.

SUBSURFACE PUMP SELECTION


Selecting a subsurface pump for a beam pumping system is a matter of estimating the pump
displacement that corresponds to a desired production rate, and then determining the optimal
combination of stroke length, pump speed and plunger diameter for this displacement. Once the
pump is sized, we can consider what type of pump is most appropriate for the given set of
operating conditions.

PUMP DISPLACEMENT REQUIREMENTS


A subsurface pump displaces a volume defined by its stroke length, pump speed (strokes per
minute), plunger diameter and volumetric pump efficiency:
PD = 0.1166 S p N D 2 E p

(1)

where PD = pump displacement at 100 percent volumetric efficiency, B/D


S p = effective plunger stroke length, inches
N = pumping speed, strokes/min
D = plunger diameter, inches
Ep = volumetric pump efficiency. This quantity is usually less than 1.0 for the following
reasons:

Leakage of fluid around the plungerThe volume of fluid that slips down around the plunger
during a pump stroke, and thus is not actually displaced, is known as slippage.
Foaming of the fluid within the pumpWhen this occurs, the pump becomes less efficient
because it not only displaces fluid but also compresses the gas phase of the foam.
Shrinkage of the fluidThe fluid pressure and temperature decrease it is produced to the
surface. This causes gas to come out of solution and the volume of the produced liquid to
shrink by a factor corresponding to the formation volume factor at the pump depth.

Local operating conditions determine the pump efficiency, which is typically in the range of 70 to
80 percent.

STROKE EFFICIENCY (ES)


The effective stroke length (Sp) in Equation 1 is the stroke length at the pump. Because of rod
stretch and contraction, acceleration and inertial effects, Sp will be considerably shorter than the
polished rod stroke length measured at the surface.
For our initial pump sizing determination, we may define the stroke efficiency (E s):

(2)

This quantity is typically on the order of 0.75 to 0.85. A more precise determination of Sp is
performed as part of the detailed rod pump system design.

ACTUAL PUMP REQUIREMENTS


Taking into account the stroke efficiency, we can modify Equation 1 to determine the pump
requirements for a desired surface production rate:

(3)
where q
D
S
N
Ep
Es

=
=
=
=
=
=

surface production rate, B/D


pump diameter, inches
stroke length at surface, inches
pump speed, stroke/minute
pump efficiency, fraction
stroke efficiency, fraction

Example:
The pumping unit currently in place at a well has a stroke length of 86 inches. What minimum
plunger diameter is required to attain a production rate of 300 STB/D at a pump speed of 14
strokes per minute, assuming pump and stroke efficiencies of 80 percent?
Solution:

For the conditions of this example, a 2-inch plunger would be more than adequate for a 300
STB/D rate. But many other combinations of S, N and D could be used to attain this rate. The
challenge in sizing a subsurface pump is to determine which of these combinations is optimal.

PUMP SIZING
Pump sizing entails placing certain limits on plunger diameter, stroke length and pump speed. For
example, if the plunger diameter is too large, it may impose unnecessarily high stresses on the
sucker rods and surface equipment. If it is too small, it will require high pump speeds to achieve
the necessary production rate, resulting in higher peak loads on the equipment. Therefore, for a
desired production rate, we must find an optimal combination of plunger diameter, stroke length
and pump speed.

PLUNGER SIZE
We begin the selection procedure by determining the optimal pump plunger size for a desired
surface production rate. This is done in Table 1 for well depths of up to 8000 feet, assuming a
pump efficiency of 80 percent (the table values are based on an internal report of the Bethlehem
Steel Company). In this case, for example, a plunger diameter of 1.25 inches would be
recommended for a surface production rate of 100 barrels per day at a pump depth of 5000 ft.
Table 1: Optimal Pump Plunger Diameters (inches)
(After Bethlehem Steel Co. Sucker Rod Handbook (1958) and Brown (1980))
Surface Production, B/D at Ep = 80%
(for polished rod stroke lengths of 74 inches or less)
Net
Lift, ft

100

200

300

400

500

600

2000

1.25 -

1.50 -

1.75 -

2.00 -

2.25-

2.50 -

1.50

1.75

2.00

2.25

2.50

2.75

3000

1.25 1.50

1.50 1.75

1.75 2.00

2.00 2.25

2.252.50

2.25 2.50

4000

1.25

1.50 1.75

1.75 2.00

2.00 2.25

2.00 2.25

2.25

5000

1.25

1.50 1.75

1.75 2.00

1.75 2.00

2.00 2.25

2.25

6000

1.25

1.25 1.50

1.50 1.75

1.75

7000

1.125 1.25

1.25 1.50

8000

1.125 1.25

Having established an optimal range of plunger diameters, we can consult the pump manufacturer
to select the appropriate pump type and production tubing diameter. Table 2 shows a typical
manufacturers recommendation.
Table 2: Recommended Tubing Diameters for Various Pump Bores
(After Axelson, Inc., 1980)
Pump Type
RH
(Rod, Heavy Wall)

RW
(Rod, Thin Wall)

TH
(Tubing, Heavy Wall)

Tubing size, inches

Pump bore, inches

1.50

1.0625

2.00

1.25

2.50

1.25 to 1.75

3.00

2.25

1.25

0.875

1.50

1.25

2.00

1.50

2.50

2.00

3.00

2.50

2.00

1.75

3.00

2.25

3.50

2.75

STROKE LENGTH AND PUMP SPEED


The optimal stroke length-pump speed (SN) combination is based on establishing a maximum
practical limit below which the rods have sufficient time to free-fall through the fluid on the
downstroke. (Figure 1) illustrates this limit graphically for the example of a conventional pumping
unit (such data are available from the manufacturer for various pumping unit types).

Figure 1

In this figure, the stroke length, in inches, is plotted on the vertical scale, and the maximum
practical limit of stroke length and pump speed is indicated by the blue line. During the
downstroke, the rods should be allowed to free-fall to avoid excessive stress on the polished rod
clamp and hanger bar. Therefore, we must select a stroke length and pump speed that fall below
this limit.
Assume, for example, the following conditions:

Desired surface production rate = 400 STB/D,


Pump efficiency = 80 percent
Plunger diameter = 2 inches.

Stroke efficiency = 85 percent.

By rearranging Equation 2 and substituting the specific values above, we obtain a relationship for
SN:
= 1262 SPM

This value is shown in the figure. Any combination of stroke length and pump speed that falls
along or to the left of this line will satisfy this specific pump design. The point where this line
intersects the maximum practical limit line suggested by the manufacturers is at a stroke length of
about 45 inches and a pump speed of about 28 strokes/min. (This pump speed is higher than
would generally be seen in the field). As a first approximation in designing the balance of the rod
pump system, use a pump design specification of D = 2 inches, S = 45 inches and N = 28 SPM.
Note that the above design procedure is not absolute. A number of wells have operated very
efficiently at pump speeds above the designated maximum. Local experience will provide the
practical knowledge needed to modify design parameters.

ROD STRING DESIGN


The API has established standard specifications for the fabrication of sucker rods, including pony
rods, polished rods, couplings, and sub-couplings (API Spec 11B, 1998). This publication shows
that sucker rods are usually made in 25 ft lengths, except in California where they come in 30 ft
lengths. The six standard rod diameters are 1/2, 5/8, 3/4, 7/8, 1, and 1 1/8 inch.

MATERIALS
Steel sucker rods normally have an iron composition of greater than 90 percent. Other elements
are added to the steel alloys to increase hardness, reduce oxidation, and combat corrosion. For
example (Weatherford, 2005):

API Grade C rods are designed for light-to-medium load applications in non-corrosive wells;
they have minimum and maximum tensile strengths of 90,000 and 115,000 psi, respectively.

API Grade D rods are designed for heavy loads in non-corrosive or inhibited environments,
and have tensile strength limits of 115,000 and 140,000 psi.

API Grade K rods are fabricated from AISI 4623 nickel-molybdenum-alloy steel, and are
designed for light-to-medium load applications in corrosive wells.

API Grade KD63 rods are fabricated from AISI 4720 nickel-chromium-molybdenum-alloy
steel, and are designed for heavy load applications in effectively corrosive wells.

One standard operating practice is to run grade D rods except where severe corrosion or H 2S
embrittlement may be a problem. Higher tensile strength rods are available for special applications
(e.g., Weatherford S-88 and EL, Norris 97, etc.). Such rods must be handled with care because
they may lose their high strength properties if damaged.
Fiberglass sucker rods were introduced in the early 1970s. Initially, their main area of
application was in corrosive wells, but it later became clear that fiberglass rods offer additional
benefits. Their lighter weight (about 70% less than steel rods) reduces the load on the pumping
unit system, allowing for higher production without changing the pumping unit. In addition, when
combined in the proper proportions with steel rods, fiberglass rods can produce longer pump
strokes than stiffer, all-steel rod strings. At the same time, fiberglass rods are more expensive than

steel rods. They also increase the complexity of the installation design, and they require a greater
degree of care in the field (Gibbs, 1991).

ROD STRESS CALCULATIONS


It is important to not exceed the maximum allowable rod stress. To calculate the maximum rod
stress, we may use the Modified Goodman equation (or, as seen in many publications, the
Modified Goodman diagram):

(1)

S A= (0.25 T + 0.5625 Smin) x SF


where: SA = maximum allowable rod stress, psi
T
= minimum tensile strength, psi
Smin = minimum rod stress, psi
SF = service factor

The specifications for tensile strength, T, have been calculated for each API rod grade. The
minimum rod stress, Smin, is measured directly or estimated for the proposed application; it is equal
to the minimum polished rod load divided by the cross-sectional area of the rod.
The service factor, SF, depends on the environment in which the rods will be placed (Table 1). SF
will be 1.0 for API grade C and D rods, when used in non-corrosive environments, but less than
1.0, for saltwater and H 2S environments. Note that the service factor for corrosive environments is
lower for grade C rods than it is for grade D rods.
Table 1: Service Factors for Sucker Rods (after Brown, 1980)
Service

API GradeC Rods

API GradeD Rods

Non-corrosive

1.00

1.00

Salt water

0.65

0.90

Hydrogen sulfide

0.50

0.70

Example: Use the Modified Goodman equation to calculate the maximum rod stress for a grade C
rod operating in a saltwater environment;

Smin =10,000 psi


the minimum tensile strength, T, = 90,000 psi
the service factor, SF, in saltwater for grade C rods is 0.65.

Substituting these values into Equation 3 yields:


S A = (0.25 x 90,000 + 0.5625 x 10,000)(0.65) = 18,281 psi.
This calculated maximum rod stress should not be exceeded during pumping operations.
In general, maximum allowable rod stresses under working conditions should not be higher than
30,000 to 40,000 psi. High tensile-strength rods are rated at 40,000 to 50,000 psi for use in noncorrosive environments.

TAPERED ROD STRINGS


The most common design for a rod string longer than about 3,500 feet is a tapered design, using
combinations of rods with different diameters, and installed with the largest diameter rod at the top
and the smallest diameter rod at the bottom. API design procedures for tapered strings are based
on a maximum of four different rod diameters, although commercially available software packages
for rod system design can accommodate more sizes.

One fundamental design criterion used for a tapered rod string is that the string should have
approximately the same unit stress in the top rod of each section. That is, the rod loads in the top
rod of each section divided by their respective cross-sectional areas should be equal. This
provides a safe design even if some corrosion pitting occurs.
In order to select the appropriate tapered rod string for a specific well, refer to API Recommended
Practice 11L (1988; reaffirmed January 2000).
Data from one of these tables is reproduced in Table 2. The sample information given is for a 5/81/2 tapered rod string.
Table 2: Rod and Pump Data for 5/8-inch - 1/2 inch Tapered Rod String
(After API RP 11L)
Rod
No.

Plunger
Diameter
(D) inches

Rod Weight
(Wr), lb. per
foot

Elastic
Constant
(Er), 106
in/lb-ft

Frequency
factor (Fe)

Rod string, % of each


size

54

1.06

0.908

1.668

1.138

44.6

55.4

54

1.25

0.929

1.633

1.140

49.5

50.5

54

1.50

0.957

1.584

1.137

56.4

43.6

54

1.75

0.990

1.525

1.122

64.6

35.4

54

2.00

1.027

1.460

1.095

73.7

26.3

54

2.25

1.067

1.391

1.061

83.4

16.6

54

2.50

1.108

1.318

1.023

93.5

6.5

5/8 inch

1/2 inch

The first column in Table 2 is the API rod number designation, given in a shorthand form where
5 indicates 5/8 inch, and 4 indicates 4/8, or 1/2 inch. The second column is the pump plunger
diameter. The third, fourth and fifth columns refer to certain rod constants, while the final two
columns show the recommended percentages of each diameter rod to be used in a tapered rod
string for this specific rod number designation.
For example, for a plunger diameter of 1.06 inches, the right-hand columns show that 44.6% of
the string should consist of 5/8-inch rods and 55.4% of 1/2-inch rods. For a 5000-foot well, this
would correspond to:

0.446 x 5000 = 2,230 ft of 5/8-inch rods on top, and


0.554 x 5000 = 2770 ft of 1/2-inch rods on bottom.

Because rods generally come in 25 ft lengths, these two numbers would be adjusted to 2,225 and
2,775 ft, respectively.
Using the data for a 3/4-5/8-inch (75) tapered string as given in Table 3, a 1.06-inch plunger
would require 27% of the rod string to have 7/8-inch rods, 27.4% to have 3/4-inch rods, and 45.6%
to have 5/8-inch rods. For the same 5000-foot well as considered above, this would yield a design
made up of

0.27 x 5000 = 1350 feet of 7/8-inch rods at the top,


0.274 x 5000 =1375 feet of 3/4-inch rods in the middle section, and finally,
0.456 x 5000 = 2275 feet of 5/8-inch rods on the bottom.

Table 3: Rod and Pump Data for 3/4-inch - 5/8 inch Tapered Rod String
(After API RP 11L)
Rod
No.

Plunger
Diameter
(D)
inches

Rod Weight
(Wr), lb. per
foot

Elastic
Constant
(Er), 106
in/lb-ft

Frequency
factor (Fe)

75

1.06

1.566

0.997

75

1.25

1.604

75

1.50

75

Rod string, % of each


size
7/8
inch

3/4
inch

5/8
inch

1.191

27.0

27.4

45.6

0.973

1.193

29.4

29.8

40.8

1.664

0.935

1.189

33.3

33.3

33.3

1.75

1.732

0.892

1.174

37.8

37.0

25.1

75

2.00

1.803

0.847

1.151

42.4

41.3

16.3

75

2.25

1.875

0.801

1.121

46.9

45.8

7.2

Designing a pumping system using the API procedure is an iterative process, including the design
of sucker rod strings. First, select a tapered rod string and calculate whether the maximum rod
stress exceeds the maximum limits for the class of rods used. If it does, then it is necessary to
select other rod number designations until a string is found that will satisfy the design limitation.
Using the API design calculations, one of these tapered strings might meet the design limitations.
If not, a heavier or lighter rod string should be designed.

ROD LOADS DURING PUMPING


Once the basic properties of the sucker rods and the design parameters of the rod string are
established, the next step in designing the string is to calculate the rod loads during pumping
operations. This information is used to size the pumping unit.
Consider the reciprocating motion of the polished rod and the actions of the traveling and standing
valves during the pumping cycle. Throughout this cycle, the load on the sucker rod string varies
continuously.

The maximum load occurs shortly after the beginning of the upstroke, when the traveling
valve closes. The polished rod must carry the full weight of the fluids, the rods, and the
added inertial effects that occur as the motion of the rods is reversed. This is known as the
Peak Polished Rod Load, or PPRL.
The minimum load occurs shortly after the beginning of the downstroke, as the traveling valve
opens. At that point, the rods no longer carry the fluid load and the inertial effects are
reversed, thereby reducing the total rod load below the weight of the rods in the produced
fluids. This is known as the Minimum Polished Rod Load, or MPRL.

To calculate the maximum and minimum rod loads, we begin with the PPRL. This is equal to
PPRL = W - Br + Fo + I
Where W = weight of rods in air, lb
Br = buoyancy of rods in fluids, lb
Fo = weight of fluid column, lb
I = effects of inertial and acceleration forces, lb.
Consider, for example, the following tapered sucker rod string:

1350 ft of 7/8-inchrods

(2)

1375 ft of 3/4-inch rods


2275 ft of 5/8-inch rods

The average density of the hole fluid is 55 lb/cu ft ( = 0.88).


The average unit weight of the rods in air is given as 1.566 lb/ft. Thus, the total weight of the rods
in air is simply equal to:
W = (5000 x 1.566) = 7830 lb
The second term, the buoyancy of rods in the produced fluids, is equal to the weight of the fluids
that the rods displace. It is calculated as follows:

(3)
Br = 0.128 W

or

where = specific gravity of the produced fluid, dimensionless


W = weight of rods in air, lb
62.4 = density of water, lb/cu ft
488 = density of steel, lb/cu ft.
Using Equation 3,
Br = 0.128 W = (0.128)(0.88)(7830) = 882 lb.
The weight of the rods in fluid, Wrf, is equal to the weight of the rods in air minus the buoyancy
effect of the rods, or:
Wrf = W - Br = 7830 - 882 = 6948 lb.
The weight of the fluid column supported by the net plunger area, Fo, is equal to the density of the
produced fluid multiplied by the net plunger area, multiplied by the height of the static fluid level.
The net plunger area is equal to the cross-sectional area of the plunger minus the area of the
rods. The suggested API procedure, however, disregards the area of the rods in these
calculations.
Fo = x A x H

(4a)

where = density of fluid, lb/cu ft


A = area of plunger, sq ft
H = fluid level, ft
Substituting = 62.4 , and A = D2/(4 x 144), where D is the plunger diameter in inches:
F0 = 0.34 D2H

(4b)

Where is 0.88, the plunger diameter is 1.06 inches and the static fluid column height is assumed
equal to the pump depth of 5000 ft:
Fo = (0.34)(0.88)(l.062)(5,000) = 1681 lb.
This is the weight of the fluid column.
Returning to Equation 2 and substituting the calculated values yields a PPRL value of 8629 lb plus
the effects of the inertial and acceleration forces, which are still unknown. Due to the complex
behavior of the polished rod, predicting the effects of these forces is the most difficult part of the
analysis. The rod string stretches and contracts in response to a cyclic motion (not a simple
harmonic) and to motions of the crank and pitman that are different for each type of pumping unit.
In addition, because of the elasticity of the sucker rod system, stress waves run up and down the
rod in response to the various forces that affect the rod.

Several methods may be used to estimate the effect of inertial and acceleration forces in the
polished rod loads.

One method is to make simplifying assumptions about the behavior of the rod and the pump
motion; however, this method is not used often today.

Another method is to calculate rod loads by solving the nonlinear partial differential equations
that represent the behavior of the rod string. Gibbs, originally on his own and later with
Neely at Shell Oil Co., solved these equations in an effort to analyze pump operation rather
than pump design (Gibbs, 1963; and Gibbs and Neely, 1966). Their work uses known values
of peak polished rod loads at the surface to calculate the behavior of the pump in the
subsurface.
Yet another method is the API design procedure that uses empirical correlations derived from
repeated observations of pump behavior. This research was conducted at the Midwest
Research Institute where computer simulations of rod-pumping systems were performed. By
simulating a wide range of pumping conditions, the Institute was able to develop correlations
for predicting polished rod loads.

Both analytical solution methods and the API procedure are incorporated into many commercially
available software programs for rod pump system design.

API SYSTEM DESIGN PROCEDURE


Designing a rod pumping system involves complex calculations of the dynamic relationships
between production rates and stresses at various points in the system. For many years, the
standard for such design calculations has been the API Recommended Practice for Design
Calculations for Sucker Rod Pumping Systems (RP 11L). First published in 1966, updated in 1988
and reaffirmed in January 2000, this recommended practice involves the use of dimensionless
variable correlations for optimizing design parameters.
Although API-RP 11L has been supplemented and in some cases supplanted by PC-based
programs that are quicker to use and more general in their application, it is discussed here to
provide some insight into the considerations involved in rod pump system design.

ASSUMPTIONS AND LIMITATIONS


The following assumptions are incorporated into API RP 11L:

Applies to conventional pumping unit motion.


Pumping unit employs medium-slip prime mover.
Steel rod stringin the case of tapered strings, the rods become smaller with depth.
Negligible friction at the stuffing box and in the pump.
Pump completely liquid-filled (no gas interference or fluid pound).
Anchored tubing (a correction formula is included to approximate effects of unanchored
tubing).
Pumping unit in balance.

DIMENSIONLESS VARIABLES
The experimental work conducted at the Midwest Research Institute (Kansas City, Missouri, 1964)
on sucker rod pumping, later adopted by the API, consisted of numerous analog computer
simulations over a wide range of pump operating conditions. The results are correlated based on
two dimensionless variables: dimensionless pump speed and dimensionless rod stretch.
The dimensionless pump speed variable may have two forms: N/No or N/No .

N/No , is the pump speed in strokes per minute (N ), divided by the natural frequency of the
string (No ).

N/N'o is equal to (N/No )/ Fc , where Fc is the frequency factor of the rod string. Fc is equal to
1.0 for non-tapered rod strings, and is greater than 1.0 for tapered rod strings. Table 4.1 of
API RP 11L lists Fc values for various tapered rod strings.
If a rod string is non-tapered, then N/No = N/N'o.

These dimensionless variables are calculated as follows:

Where A = speed of sound in steel rods (approximately 980,000 ft/min)


L = rod length, ft
Substituting this value of No into the dimensionless variable,

(1)
and

(2)
Thus, for a given or assumed pump speed, rod string length and frequency factor, we may
calculate a value for each of these dimensionless variables.
Before making this calculation, consider the second dimensionless variable, the dimensionless rod
stretch, defined as
Fo/Skr.
where
Fo = the static fluid load in pounds
S = the polished rod stroke length, in inches
k r = the spring constant of the rod string.
k r can be derived from Er (the reciprocal of kr multiplied by the rod length, L), which is also listed
in Table 4.1 of API RP 11L (Figure 1: Excerpt from Table 4.1 of API RP 11L). Skr represents the
load in pounds required to stretch the rod string the length of the polished rod.

Figure 1

The dimensionless rod stretch term represents the rod stretch caused by the static fluid load; it is
given as a fraction of the polished rod stroke length. (Fo/Skr = rod stretch as a fraction of polished
rod stroke lengthfor example, if Fo/Skr = 1, then the rod string will stretch an amount equal to
the length of the polished rod. The value of this dimensionless variable may be easily calculated
for a known or assumed polished rod stroke length.
On a practical note, observations have determined that under-travel and over-travel of the pump
plunger can be prevented if N/No' is less than 0.35, and Fo/Skr is less than 0.50.
From the work performed at the Midwest Research Institute, five correlations were developed
these are shown in API RP 11L as Figures 4.1, 4.2, 4.3, 4.4 and 4.5. These five correlations,
based on the two dimensionless variables for their presentation, are used to systematically
develop the API design procedure for a pumping system.
The first three correlations affect the rod and downhole pump systemspecifically, the plunger
stroke length, the peak polished rod load, and the minimum polished rod load. The last two
correlations are used to calculate peak torque and polished rod horsepower.

DESIGN EXAMPLE:
To demonstrate the use of these correlations consider a well that is to be pumped at a rate of 100
B/D. The well conditions are:

Fluid level (H) and pump depth (L) = 5000 ft


Tubing diameter = 1.9 inches.
Tubing is anchored
A tapered sucker rod string (API 75), consisting of 1350 ft of 7/8-inch rods, 1375 ft of 3/4-inch
rods, and 2275 ft of 5/8-inch rods, is calculated to be sufficient for this design. These grade
D rods have a maximum allowable rod stress of 34000 psi.
Plunger diameter = 1.06 in.
For preliminary design, pump speed (N) = 16 strokes/min
Polished rod stroke length (S) = 64 inches.

Specific gravity () of the pumped fluid = 0.88.

Substituting these data into the equation for pump displacement, using the polished rod stroke
length, S, as an estimate of the pump stroke length, and assuming a pump efficiency of 100
percent, we obtain:
PD = 0.1166 Sp N D 2 Ep = (0.1166)(64)(16)(1.06) 2 =135 B/D

(3)

This quick calculation gives us a pump displacement of about 135 B/D, which should satisfy the
wells IPR-calculated objective of 100 B/D.

PLUNGER STROKE CORRELATION


The first correlation allows us to calculate the plunger stroke length ( Figure 2: Plunger stroke
correlation, after API RP 11L).

Figure 2

The dimensionless pump speed, N/No , is shown on the horizontal axis.


The bottomhole plunger stroke length divided by the polished rod stroke length (S p/S ) is
plotted on the vertical axis. This ratio reflects the effect of rod stretch on the effective plunger
stroke length.
The plot shows a series of curves, each for a different value of the dimensionless rod stretch
(Fo /Skr). Because the values of N/No and Fo /Skr are knownor can be calculated from the
given datawe may use this correlation to calculate a value for Sp /S .

For the example problem, N/No is calculated as follows:


N/No = NL / 245,000 = (16 * 5,000) / 245,000 = 0.326

For the selected tapered rod string, F c is equal to 1.191, and so:
N/ N'o = N / (No Fc ) = 0.326 / 1.191
N/ N'o = 0.274
The dimensionless rod stretch is equal to Fo/Skr.
Fo, the weight of the fluid supported by the rods, is
F0 = 0.34 D2H

(4)

where = specific gravity of fluid of fluid, lb/cu ft =


D = plunger diameter, inches
H = fluid level, ft
Where is 0.88, the plunger diameter is 1.06 inches and the static fluid column height is assumed
equal to the pump depth of 5000 ft:
Fo = (0.34)(0.88)(l.062)(5,000) = 1681 lb.
The polished rod stroke length (S) is 64 inches, and the reciprocal of k r is equal to the elastic
constant, E r for the example rod string multiplied by its length, L (i.e., 1/k r = E rL). From API
manual RP 11L , we obtain, E r.
E r = 0.997 x 10 -6 in/lb ft
and substitute the value into the relationship for the dimensionless term:
Fo/Skr = (1.681 * 0.997 * 10-6 * 5000) / 64
Fo/Skr = 0.13
Note that N/N'o < 0.35, and Fo / Skr < 0.50. This puts them within the practical limits for preventing
under-travel and over-travel of the pump plunger mentioned above.
Referring back to (Figure 2), we see that with known values for the two dimensionless variables,
we arrive at a value of S p / S = 0.98. This means that if the tubing is anchored, as in this example,
the bottomhole plunger stroke length is equal to 98 percent of the polished rod stroke length at the
surface. Specifically, for a polished rod stroke length of 64 inches, the bottomhole stroke length will
be 62.7 inches.
If the tubing is not anchored, the plunger stroke length corrected to account for the tubing
contraction during the upstroke. This term is found by the following formula:
Tubing contraction = Et Fo L

(5)

Where Et = coefficient of elasticity for the tubing, in/lb ft


Fo = weight of fluid on rods, 1b
L = tubing length, ft.
For 1.9-inch tubing, Et has a value of 0.5 x 10 -6 in/lb ft. With Fo equal to 1681 lb, and L equal to
5000 ft, the tubing contraction in unanchored tubing, is:
Tubing Contraction = (0.5 x 10 -6)(1681)(5000) = 4.2 in.
So, in this example, the net plunger stroke length would be (62.7 - 4.2) = 58.5 inches.
We can now calculate the subsurface displacement. Assuming that the tubing is anchored, the
pump displacement is:
PD = 0.1166SpND 2 = (0.1166)(62.7)(16)(1.06) 2= 131 B/D.
This more than meets the example expected inflow rate of 100 B/D. In fact, it allows more
downtime per day for maintenance. However, had the pump displacement rate not met the
production objective, it would be necessary to revise the assumed pump data and repeat the
calculations. Note that to determine the production rate at the surface, we must incorporate the
pump efficiency (Ep) into the calculation.

PEAK POLISHED ROD LOAD CORRELATION


The second API correlation can be used to calculate the peak polished rod load ( Figure 3: PPRL
correlation, after API RP 11L). The same dimensionless variables, whose values are known, are
plotted in the same general locations. In this case, the horizontal axis is simply N/N o . It is the
vertical axis that is changed, and is now equal to Fl/Sk r . The terms in the denominator are already
defined; their values were calculated earlier. Fl , the only unknown, is referred to as the peak
polished rod load factor.

Figure 3

The dimensionless pumping speed, N/No, is equal to 0.326. The dimensionless rod stretch, Fo/Skr,
is equal to 0.13. With these two known values, we can use ( Figure 3) to find a value of 0.37 for
Fl/Skr.

Because Skr is known, it is possible to calculate Fl. Note that Fl must be added to the weight of
rods in the produced fluids to give a value for the peak polished rod load in pounds. It thus
represents two terms: the weight of the fluid carried by the rods and the load caused by the
inertial-acceleration forces.
The peak polished load, then, is equal to:
PPRL = Wrf + Fl

(6a)

Where PPRL = peak polished load, lb


W rf = weight of the rods in fluid, lb
F1 = weight of fluids on rods plus inertial-acceleration forces, lb
Or, in terms of the correlation:

(6b)
Where PPRL = peak polished load, lb
Wrf = weight of the rods in fluid, lb
F1
= weight of fluids on rods plus inertial-acceleration forces, lb
Skr = the load in pounds required to stretch the rod string the length of the polished rod.
For this sample problem, we determine the weight of the API 75 rod string in the produced fluid to
be 6948 lb. The second term (F l/Skr) is equal to 0.37. All that is needed to calculate PPRL is the
value of Skr.
The spring constant for the rod string (k r) is equal to the reciprocal of the coefficient of elasticity of
the rod string (E r ) multiplied by the length of rods (L):
1/k r = (E r)(L)
1/kr = (0.997 x 10 -6 )(5000).
Then, taking the reciprocal, we obtain:
kr = 200.6.
With a polished rod stroke length, S, of 64 inches, Sk r is calculated :
Skr = (64)(200.6) = 12838 lb.
With this information the peak polished rod load for the example can be calculated:
PPRL = W rf + ((F l/Skr) * Skr)
PPRL = 6948 + (0.37 * 12838) = 11698 lb.
This value represents the best estimate of the rod load, and can be used for design purposes.
Under actual operating conditions, it may be different, but not significantly so.

MINIMUM POLISHED ROD LOAD CORRELATION


The third API correlation (Figure 4: MPRL correlation. After API RP 11L) involves calculating the
minimum polished rod load (MPRL). This load occurs just at the beginning of the downstroke,
when there is maximum downward inertial force, and the rods no longer carry the fluid load.

Figure 4

The two dimensionless variables are in the same relative locations, but now the term F2/Skr is on
the vertical axis. F2 is defined as the minimum polished rod load factor.
Using the information already obtained, with an N/No of 0.326, and an Fo/ Skr value of 0.13, the
point of intersection on the vertical axis is at 0.21. With the value for Skr known, we can calculate
F2 .
F2 is the weight that must be subtracted from the weight of the rods in the produced fluid to give
the MPRL. As such, it represents the inertial and acceleration force at the point of minimum load.
Its calculation proceeds directly parallel to that for the peak polished rod load:

(7a)

MPRL = W rf -F2
MPRL = W rf - (F2/Skr * Skr )

(7b)

MPRL = 6948 - (0.21 * 12,838)


MPRL = 4252 lb.
This MPRL of 4252 lb. is used for design purposes; the actual value of F2 after installation may be
different.

SELECTION OF COUNTERWEIGHTS
Counterweights are added to a pumping unit to provide a counterbalance effect at the polished
rod. We can appreciate their importance by considering what would happen if this counterbalance
was not present. On the upstroke, the prime mover would have to do an enormous amount of
work. An excessive torque would be exerted on the gear reducer, requiring a larger prime mover
and a more substantial gear reducer. But on the downstroke, with the polished rod load
substantially reduced and the force of gravity pulling the rods and plunger downward, the prime
mover would have little work to do. In short, it would be practically impossible to design a system
to efficiently handle such load variations.
Counterweights serve to reduce these load variations by moving downward during the upstroke
(thus helping to lift the rod string) and upward during the downstroke (thus keeping the rod string
from falling too quickly). This reduces the size requirements for the prime mover and provides a
more even load to the gearbox.
For design purposes, we can estimate of an ideal counterbalance that establishes an
approximately equal balance between (1) the upstroke and downstroke work done by the prime
mover and (2) the net torque exerted on the gear reducer during each half of the pumping cycle.
This balance should be the design objective, and is determined by calculating the maximum and
the minimum loads on the polished rod during the pumping cycle and then calculating their mean
value:
Mean value = W rf + 0.5 Fo

(8)

Where Wrf = weight of rods in fluid, lb.


Fo = weight of fluid imposed on rods, lb.
The required counterbalance effect, CBE, should be approximately equal to the mean load. For
API design purposes, however, the design value of the counterbalance effect for the conventional
pumping unit is slightly larger than this amountin fact, it is 1.06 times larger than the mean load.
CBE = (1.06) (Wrf + 0.5 Fo).

(9)

Substituting the data from the sample problem into this equation, we obtain the total required
counterbalance effect:

This CBE is required at the polished rod; it is transmitted there through the counterweights acting
on the connecting pumping unit members. To determine the size of counterweights needed, a
force balance around the various contributing structural members must be calculated that includes
the effect of the units geometry and its contribution to the CBE. The pumping unit manufacturer
provides this information. For a given pumping unit, pump stroke length, and desired CBE, the
manufacturer will specify the types of counterweights to be used and where they should be placed
on the unit.
The actual CBE needed for a pumping unit is determined by field measurements of the peak and
minimum polished rod loads. These measurements will dictate to what position the counterweights
must be adjusted or, for an air-balanced unit, by how much the pressure in the cylinder must be
changed.

PEAK TORQUE CORRELATION

Consider the torque that is imposed on the crank by the low speed shaft of the gear reducer
during the pumping motion. One important system design objective is to be sure that the peak
torque does not exceed the manufacturers limits.
Remember that torque is equal to the force acting at right angles to a lever arm, multiplied by the
length of the arm, and that torque tends to produce rotation at the point of connection. The torque
of a pumping unit is provided by the prime mover, and is applied to the crank by the low speed
shaft of the gear reducer that, in turn, causes the pump to operate ( Figure 5: Torque generated by
pumping unit). This torque must be sufficient to cause the pump to operate in a continuous
manner under normal operating conditions.

Figure 5

The net torque at the gear reducer is equal to the torque caused by the well loads acting on the
polished rod minus the torque caused by the counterweights acting along the crank. Because the
magnitudes of these two terms change during the pumping cycle, the peak torque is defined as
the maximum net torque that occurs during the pumping cycle.
(Figure 6) shows the net torque imposed on a typical conventional beam pumping unit.

Figure 6

The dashed curves show the rod load and counterweight components of the torque during one full
cycle, beginning with the upstroke. Notice that the rod load during the upstroke is positive and the
counterweight torque is negative. The maximum and minimum values occur about midway through
the upstroke. As one value decreases, the other increases. The values become equal and very
close to zero at the top of the upstroke, and reach maximum and minimum points midway through
the downstroke. The solid curve shows the net effect of the two componentsit reaches a
maximum value (peak torque) midway through the up and down portions of the stroke. The gear
reducer, then, must be designed to accommodate this peak net torque by an appropriate safety
margin. Manufacturers specify the peak torque for each pumping unit. Note that the above
representation is for a conventional unit and will be different for other types of units.
The peak torque of a pumping unit is calculated by using two additional API correlations. Peak
torque, PT, is given by the multiple of four terms:

(10)
Where PT = peak torque, lb-in
T
= torque, lb-in
Ta = torque adjustment factor, fraction
Skr = load required to stretch rod string the length of the polished rod, lb.
S
= polished rod stroke length, in.
Notice that the second and third terms in the example problem are already known. It is the first
and the fourth terms that must be obtained from the two new correlations.
The peak torque correlation is shown in (Figure 7) (Peak torque correlation, after API RP 11L).
The first term in on the right-hand side of Equation 10 is shown on the vertical axis. The two
familiar dimensionless variables (N/No and Fo/Skr) are shown in the same relative positions as they
were in earlier correlations.

Figure 7

Starting from N/No =0.326 on the horizontal axis, and moving vertically upward to a value of Fo/Sk r
= 0.13, we determine that the corresponding value of (2T/S2kr ) is 0.32.
Thus, for our example design, Equation 10 has the following form:
PT = (0.32) (12838) (64/2) Ta =131461 T a.
A single unknown, Ta , remains to be determined. Where (W rf / Sk r ) = 0.3, then Ta = 1.0, and
no further adjustments to the equation are necessary. Where (W rf / Sk r 0.3), then Ta differs
from 1.0. In order to calculate its value under these conditions, turn to the fourth API correlation
shown in (Figure 8) (Peak torque correlation, after API RP 11L, continued) .

Figure 8

Here, the familiar dimensionless variables are found on the vertical and horizontal axis. In this
case, the horizontal axis is N/No . Within the chart are percentage values ranging from positive
values on the outer rim to negative values near the center. For specific values of the
dimensionless variables, this chart is used to find the appropriate percentage. The value of this
percentage is then used to calculate the torque adjustment constant, Ta , using the following
equation:

(11)

Where Ta
X
W rf
Sk r

= torque adjustment factor, fraction


= percentage correction
= weight of rods in fluid, lb.
= load required to stretch rod string the length of the polished rod, lb

Using the specific values for the dimensionless variables that have been used previously (i.e.,
0.274 on the horizontal axis and 0.13 on the vertical axis) the percentage correction is found to be
equal to about +2.5% (Figure 8 , above).
Substituting this value into the equation,:

=1.06
When this value is included into the relationship for peak torque, PT:
PT = (131,465)(1.06) =

= 139,353 in/lb.

This is the expected peak torque of the example pumping unit. The shaft of the gear reducer
should be designed to withstand a maximum torque of at least this value. Later, this design
specification will be introduced into the pump unit designation.
The gear reducer, acting through the V-belts, transmits the power and the rotating speed of the
prime mover into the power and rotating speed needed by the pumping unit. Normally there are
both high-speed and low-speed gears, and a careful selection of the gear reduction system is a
necessary part of the manufacturers overall design of the pumping unit.

POLISHED ROD HORSEPOWER CORRELATION


The API procedure for calculating the polished rod horsepower, PRHP, for the conventional unit is
given by the following relationship:
PRHP = (F3 /Skr) Skr SN (2.53 10-6)

(12)

where F3
= PRHP factor, lb and, as defined above
Sk r = load required to stretch rod string the length of the polished rod, lb
S
= stroke length at surface, inches
N
= pump speed, stroke/minute
A value for F3 may be found by using the fifth API correlation, as shown in ( Figure 9) (PRHP
correlation, after API RP 11L).

Figure 9

Notice this figures similarity to the other API correlations. The two familiar dimensionless variables
appear in the same relative positions, and the new variable appears on the vertical axis. This
variable is the first term in the calculation for the polished rod horsepower. To find its specific
value, enter the horizontal axis the previously determined N/N o value of 0.326, and move
vertically upward to a F o /Sk r value of 0.13. These correspond to F3 /Sk r = 0.225.
Substituting this value plus the known value of other terms into Equation 12 yields:
PRHP = (0.223)(12,838)(64)(l6)(2.53 x 10 -6) = 7.48 HP
Knowing the PRHP, we can determine the nameplate horsepower (HPnp), which is the minimum
horsepower required to run the unit:
HPnp = (PRHP)(CLF) / (E surf )

(13)

where PRHP = polished rod horsepower,


CLF = cyclic load factor
E surf = surface efficiency of pumping unit, fraction
The cyclic load factor (CLF) represents the additional reserve power needed to handle the cyclic
beam-pumping load. For low peak torques and uniform torque ranges, the cyclic load factor is low
when peak torques are low and torque ranges are uniform. If, on the other hand, the peak torques
are high and the ranges are non-uniform, the cyclic load factor is high.
The prime mover normally recommended for beam pumping units is the NEMA Design D electric
motor. For this unit, the CLF is about 1.375. For NEMA C electric motors and multi-cylinder
engines, the CLF will be about 1.897. These CLF values are valid for conventional and air
balanced units, but will differ for MARK II units, as shown in Table 1.
Table 1: CLF Values for Various Pumping Unit Types
(Courtesy of Lufkin Industries, Lufkin, TX, USA)
Unit Type

Prime Mover Type

CLF
Value

Conventional
and
Air Balanced

NEMA D electric motors and slow speed


engines

1.375

NEMA C electric motors and multi-cylinder


engines

1.897

Mark II

NEMA D electric motors and slow speed


engines

1.10

NEMA C electric motors and multi-cylinder


engines

1.517

The third term in Equation 13 is the surface efficiency, Esurf. It is equal to the combined efficiencies
of the motor and the moving parts of the pumping unit, including the wire lines, structural bearings,
transmission, gears, gearbox bearing, and the V-belt.
The mechanical efficiency is usually about 90%. The motor efficiency depends on the motor, its
operating speed and load variation during a revolution of the crank. The motor manufacturer will
provide motor efficiency information. For this example, assume that it is 70 percent.
Esurf = (0.9)(0.7) = 0.63.
This information may be used to calculate the nameplate horsepower for the example problem.
Assume that a NEMA D electric motor, with a CLF of 1.375, is installed.
A number of factors, including polished rod horsepower, torque loading, system voltage, start-up
problems and motor characteristics, will influence motor size. Substituting the known values into
Equation 13 yields:
HPnp = (7.48)(1.375) / 0.63 = 16.3 HP
The nameplate horsepower for this NEMA D motor is found to be 16.3. Notice that the motor
nameplate horsepower is about two times the polished rod horsepower.
Because motors come in discrete sizes, as shown in Table 2, a 20 horsepower unit would be
ordered for this example application.
Table 2:
Recommended Motor Sizes for API Pumping Units
(Courtesy of Lufkin Industries, Lufkin, TX, USA)
Unit Size

Recommended Motor Sizes,

HP
57-109-42

7.5

10

80-133-54

7.5

10

114-169-64

7.5

10

15

160-200-74

10

15

20

228-212-86

15

22

30

320-256-100

20

30

40

456-304-120

30

40

50

640-356-144

40

50

60

There are a number of other ways to calculate the horsepower required for the prime mover.
Some have been developed by manufacturers, others by individual oil companies. Table 3 shows
examples of different nameplate horsepower calculations used by one manufacturer.

Table 3: Alternative Nameplate Horsepower Calculations


Conventional and
Air Balanced Units

Mark II Units

Slow speed engines


and high-slip electric
motors
Multi-cylinder engines
and normal-slip
electric motors

ROD STRESS CALCULATIONS


The last calculation in the design procedure is made to ensure that the rod stress is not too high.
Maximum rod stress is equal to the peak polished rod load divided by the cross-sectional area of
the top rod:
Rod Stress = PPRL / Rod Area

(14)

In this design example, the rod stress exerted on the 7/8-in top rod is (11,698 / 0.601), or 19,464
psi. This value is well below the maximum allowable stress of 34000 psi that was specified for the
Grade D rods.

POLISHED ROD DYNAMOMETERS


Rod-pumped wells must be monitored to ensure the continued efficient and economic operations
of a field. After a pumping unit has been installed in the field, dynamometer tests of pumping wells
are conducted to determine the system efficiency, and whether adjustments need to be made in
stroke length, pump speed or other operating parameters. These tests are designed to answer
such well performance questions as (Podio et. al, 2001):

Is the well pumped off?


What is the pump intake pressure?
What is the pump fillage?
What is the pump displacement?

Is the standing valve or traveling valve leaking?


What is the effective pump plunger travel? What is the current pumping speed?
What is the polished rod horsepower?
Is the gearbox overloaded?
Is the unit properly balanced?
Is the downhole gas separator operating efficiently?

A dynamometer survey measures the load forces acting on a rod string during a complete
pumping cycle and records the forces on a chart or computer display. This display is often called a
dynamometer card. The card records changes in the rod load versus rod displacement, or
changes in the rod load versus pumping time (Figure 1: Example of surface (top) and pump
(bottom) dynamometer cards. Courtesy Echometer Company, 2003).

Figure 1

During a pumping cycle, forces acting on the rod string cause changes in the rod load.
Measurements of these rod loads reflect the operation of the subsurface pump and the surface
unit.
Polished rod dynamometers record polished rod loads during the pumping cycle. Most common
dynamometers make continuous plots of load measured against the movement of the polished

rod. Typical conventional dynamometers are either mechanical or hydraulic, but electronic
dynamometers are also used. These dynamometers are briefly described below.

MECHANICAL DYNAMOMETER
A mechanical dynamometer measures rod loads by measuring the deflection of a steel ring placed
between the carrier bar and the polished-rod clamp. A recording of this ring deflection is made on
a paper dynamometer card, attached to a rotating drum driven by the vertical movement of the
polished rod. The card traces polished-rod loads against rod vertical displacement. Its major
disadvantage is that the pumping unit must be stopped before it can be installed on the polished
rod.

HYDRAULIC DYNAMOMETER
A hydraulic dynamometer (Figure 2: Courtesy Echometer Company, 2003) can be installed while
the pumping unit is running. This type of dynamometer uses a spacer installed on the polished rod
between the carrier bar and the polished-rod clamp. As the unit is pumping, two load-sensing
hydraulic pistons can be installed between the shoulder of the spacer and the carrier bar. The
polished-rod load is transferred to the hydraulic pistons. Changes in polished-rod loads affect the
hydraulic pressure. These changes in hydraulic pressure are recorded on a drum chart as
changes in rod load versus rod displacement.

Figure 2

ELECTRONIC DYNAMOMETER

The electronic dynamometer uses electronic transducers to measure well loads and rod
displacements (Figure 3: Polished rod dynamometer transducer. Courtesy Echometer Company,
2003).

Figure 3

An electronic dynamometer units main components include a load transducer (load cell), a
position transducer and the electronics that provide interfacing, signal recording, and processing.
The load cell, located between the carrier bar and the polished-rod clamp, uses a strain gauge to
measure polished rod loads. The position transducer uses a potentiometer to measure polishedrod displacement. Accelerometers determine rod displacement so that polished-rod load and
position can be recorded as a function of time. In this manner, an electronic dynamometer can
produce surface dynamometer cards and supply the information needed to construct downhole
cards.
Portable electronic dynamometer systems usually include microcomputers that combine real-time
data acquisition with easy data storage and retrieval operations, as well as performing on-line
analysis of measurements and calculating downhole cards.

DYNAMOMETER CARD INTERPRETATION


(Figure 4) shows a dynamometer card shortly after it is recorded at the surface.

Figure 4

A number of measurements must be noted on the card.

The first is C, the calibration constant in pounds per inch of card height. Each inch of height,
then, refers to some calibrated load on the polished rod. This is determined in the field.

D1 is the maximum deflection in inches.


D2 is the minimum deflection in inches.
CB is the counterbalance line, which is obtained by drawing a line through the dynamometer

card at the point that represents the static load on the polished rod when the crank arm is
horizontal. This means that the pumping unit is stopped and the load on the polished rod is
measured with the crank arm at an angle of 90 or 270 0 from the 12 oclock position.
A 1 is the lower area of the card, measured in square inchesthat is, it is the area between
the lower boundary and the zero line.
A 2 is the area within the card shape, measured in square inches.
Finally, L is the length of the card, in inches, measured along the horizontal axis from one end
of the dynamometer shape to the other.
The left-hand extreme of the card is the bottom of the stroke, and the right-hand extreme is
the top of the stroke.

With a dynamometer card for a given rod pump and the definitions above, we can make the
diagnostic calculations listed in Table 1:
Table 1: Diagnostic Calculations for Beam Pumping System

Peak polished rod load

PPRL =D 1C

Minimum polished rod load

MPRL = D 2C

Polished rod load range

C(D 1 - D 2)

Counterbalance effect

Actual CBE = CD 3

Approximate CBE

(C/L) ((A 1+(A 2/2))

Polished rod horsepower

PRHP = C(A2/L) (SN/33000 x 12)

The final term to be calculated is the peak torque that is exerted on the gear box under operating
conditions. Normally this calculation is made using a term referred to as the torque factor, TF. The
torque factor is equal to the torque that must be applied at the crankshaft to offset a one pound
weight hanging on the horsehead. Thus, if a 25 ft-pound torque must be applied at the crankshaft
to offset a one-pound weight hung on the horsehead, the torque factor is 25. The torque factor will
vary with different crank angles.
With a value for TF, the net torque calculation, which is a function of the crank angle, can be made
by using the following equation:
Net torque = TF (W-B) - M Sine q

(Equation 15)

where: W = well load for a given crank angle, pounds


B = structural imbalance of pump unit, pounds (supplied by manufacturer)
N = torque provided by the counterweights, ft-pounds
TF = torque factor
q = crank angle
Using this equation, we can look at the dynamometer card, insert the crank angle on it for a
number of different values of W and TF, and make net torque calculations using the above
equation. Note that W is a function of the crank angle and is measured from the dynamometer
card. The torque factor, TF, is supplied by the manufacturer for a given unit in 150 angles of the
crankshaft.
This net torque calculation should provide a plot similar to that shown in the dashed curve of
(Figure 5), showing the peak torque at the maximum point on the curve. It suggests that the
counterweights should be adjusted in order to even out the net torque curve. After an adjustment
of the counterweights, for example, a second curve is obtained (solid curve), showing that the unit
is now more balanced.

Figure 5

The downhole condition of the pump and its pumping action can be analyzed with a dynagraph.
With the dynamometer, those terms that affect the surface operating condition can be calculated
and necessary adjustments to the unit can be made. If the operation of the subsurface pump and
the surface unit are satisfactory, is it possible to improve production by changing the pump stroke
length, its speed or its surface control system? Generally, a unit that runs intermittently is
controlled either by a timer or with a pump-off controller. Careful adjustments in the field are
required to maximize production, without damaging the pumping unit. The information from
dynamometer tests taken both before and after changes are made in the pumping equipment will
help determine whether the adjustments have improved the operating conditions.

ROD LOAD VERSUS DISPLACEMENT


To understand the diagnostic procedures involved in monitoring a beam pumping system, we must
first understand the principles of dynamometry. Gibbs and Neely (1966) discuss polished-rod load
versus displacement for anchored and unanchored tubing. Typical plots for the unanchored and
anchored tubing case are shown in the next series of figures. These plots are developed
systematically.

UNANCHORED TUBING

(Figure 1) represents the rod load as a function of displacement for unanchored tubing. Note that
the load scale goes as high as 10,000 pounds, and that the pump displacement (in feet) has a
coordinate system that runs from a point above the top of the subsurface pump down to the
standing valve. When the plunger is at its lowest point, it carries no load; its displacement is 9 feet
from the coordinate to the ground coordinate system, but it is stationary with respect to the
stretching tubing.

Figure 1

In (Figure 2), as the traveling valve reaches point C, the entire fluid load is supported by the
rods, until it begins to transfer to the tubing at point D.

Figure 2

In (Figure 3), as the traveling valve reaches point E, the transfer of the load from the rods to the
tubing is almost complete, but the tubing continues to stretch somewhat. At point F, the entire
fluid load is now supported by the tubing, causing the tubing to stretch to its greatest length. The
plunger now descends with respect to both the ground and tubing coordinate systems. At this
point, there is no load on the traveling valve and it continues to descend until it reaches a
displacement position, indicated by point A. The cycle then repeats.

Figure 3

ANCHORED TUBING
If the tubing is anchored, it will not stretch or contract in response to the movement of the pump,
and the ideal load-displacement diagram will have the appearance shown in ( Figure 4) (Ideal
load-displacement diagram for a pump with anchored tubing).

Figure 4

PUMP VALVE DYNAMOMETERS


Efficient operation of a sucker-rod pump depends on the integrity of the standing and traveling
valves. If these valves lose their ability to hold a seal, the lifting capacity of the pump is affected.
Therefore, the valves should be tested periodically with a dynamometer.

TESTING THE STANDING VALVE


Dynamometer testing of the standing valve requires that pumping unit to be stopped near the
bottom of the downstroke. The initial polished-rod load is recorded on the dynamometer chart.
Since the fluid load is completely carried by the standing valve, the polished-rod load recorded at
the start of the test will measure only the buoyant weight of the rod string.
When the standing valve holds a good seal, the polished-rod load will remain steady, resulting in a
line on the dynamometer card that falls on the first measurement. If there is a leak in the valve
seat, the pressure immediately below the traveling valve will decrease, then slowly close. As the
test proceeds, the fluid load initially carried by the standing valve will move to the traveling valve.

This movement increases the load of the polished rod, and will be recorded on the dynamometer
chart. The rate of load increase is directly proportional to the severity of the leak in the standing
valve and it is not uncommon for a leaking standing valve to show a load increase in about 20
seconds.
This test should be repeated. If the same results are obtained, then the test indicates that the seat
of the standing valve is cut. If the valve holds during some of the tests, then a damaged valve ball
is indicated.

TESTING THE TRAVELING VALVE


Testing the traveling valve will help to distinguish whether the traveling valve or the barrel-plunger
fit is leaking. To start the test, the pumping unit is stopped near the top of the upstroke and the
initial polished-rod load on the dynamometer is recorded.
The initial recorded polished-rod load measures the buoyant weight of the rod string and the fluid
load acting on the plunger. The standing valve is open so it carries no load. If the seal in the
traveling valve is good, and additionally, the pump barrel and plunger fit is good, then no change
in polished-rod load should be observed over the test period.
A leak either in the traveling valve or between the pump barrel and plunger will allow fluid to pass
through the traveling valve. This leak will slowly force the standing valve to close. Once closed, the
fluid load transfers to the standing valve and the tubing, and this load transfer will be observed on
the dynamometer card. The initial polished-rod load (rod-string weight plus fluid load) recorded on
the card will be greater than the loads measured later in the test. The rate of load decrease is
proportional to the leak rate.
This test cannot distinguish between a leaking traveling valve and leaking caused by a worn barrel
or plunger. This test should be repeated at different points in the upstroke. Load differences
observed at different depths can help to identify the location of the leak.

MEASURING THE CBE


A dynamometer can also measure the counterbalance effect (CBE) that is experienced when a
pumping unit is overbalanced, or "weight heavy". This condition occurs when the counterweights
move down and lift the polished rod after the motor is stopped and the brake is released. If the
pumping unit is underbalanced, or "rod heavy", the polished rod will move down and lift the
counterweights after the unit is stopped.
The CBE is measured with the dynamometer in place, carrying the full polished-rod-load. After the
motor is stopped, the brake is used to stop the pumping unit with the crank angle at 90 or 270,
so that the torque exerted on the crankshaft counterweights is at a maximum. The vertical force on
the polished rod resulting from this torque is the counterbalance effect.
The CBE value measured above is useful for estimating the maximum torque.
Pumping units with non-conventional geometries have counterweights attached to arms that are
set at angles offset to the cranks. The maximum counterbalance torque is exerted when the
counterweight arms, rather than the cranks, are in a horizontal position. However, the API
recommends that CBE be determined with the cranks held horizontal during the measurement.

SYSTEM MONITORING: DYNAGRAPHS


Surface dynamometer cards cannot be used to directly measure the operations of a downhole
pump, because the surface measurements record all the static and dynamic forces acting on the
rod string. We can make a more accurate evaluation of the pump by placing a dynamometer just
above it. Measurements of rod loads recorded immediately above the pump are called dynagraph
cards, to distinguish them from surface-recorded dynamometer cards.
To illustrate the use of dynagraph cards, consider the idealized typical profiles shown in the
following figures.

In (Figure 1) , a small, gradual curvature on the dynagraph during the upstroke indicates that
a small amount of gas, perhaps coming out of solution, is being compressed during the
upstroke. As the pump begins its downward stroke, the load is not entirely transmitted to the
tubing due to a period of gas compression on the downstroke before the traveling valve
opens.

Figure 1

In (Figure 2) , the fall-off in load is sharper during the downstroke than for the gas
compression dynagraph of Figure 1. This type of response generally means that the well is
pumped-off or almost pumped-off. In this condition, the pump chamber doesnt completely
fill with fluid. As the plunger moves down and compresses the gas, the pressure below the
traveling valve is not high enough to prevent the plunger from striking the liquid interface in
the lower part of the chamber. The shock that results when the plunger hits the fluid
interface and releases the load, is called fluid pound. Often, fluid pound can be felt at the
surface, indicating that the well is pumped-off. Dynagraph cards recorded on wells with gas
compression look similar to wells that experience fluid pound, and it can be difficult to
distinguish between the two.

Figure 2

(Figure 3) is a dynagraph profile indicating low pump efficiency. The pump is gas locked and
has no fluid in its chamber. The pump is simply causing gas to compress and then expand.

Figure 3

(Figure 4) indicates a leaking traveling valve. Movement during the upstroke is required before
a full load is taken on and, because the traveling valve is leaking, the rods begin to lose load
before the pump has reached the top of the upstroke.

Figure 4

(Figure 5) indicates a leaking standing valve. The leak causes premature loading of the rods
before the beginning of the upstroke, and also delays the unloading of the rods during the
downstroke, thus the traveling valve closes during the downstroke, causing it to bear some
of the load.

Figure 5

If (Figure 4) and (Figure 5) were combined, the results would look like (Figure 6). This
dynagraph indicates a worn-out pump that is leaking at the traveling valve, at the standing
valve, and probably has excessive plunger slippage, too. This pump should be replaced.

Figure 6

SYNTHETIC DYNAGRAPHS
The dynagraph is a valuable diagnostic tool for analyzing pump performance. Unfortunately,
because it is located at the pump depth, this analysis may be difficult to obtain. Alternative
procedures include:

Using dynamometer cards of load displacements measured at the surface to interpret what is
happening in the subsurface, and

Using surface measurements and procedures (Gibbs, 1963 and Gibbs and Neely, 1966), to
develop mathematically derived dynagraphs called synthetic dynagraphs (often referred to
as pump cards).

In the latter case, Gibbs and Neely made it possible to measure loads and displacements at the
surface and then, to calculate the required synthetic dynagraph by solving the appropriate nonlinear partial differential equations.
Obtaining a synthetic dynagraph is fairly straightforward. The polished rod load versus time
(Figure 7) and the polished rod displacement versus time (Figure 8) during a pumping cycle are
measured.

Figure 7

Figure 8

These two graphs are then combined to give a polished rod load versus polished rod
displacement profile.
Using the data collected at the surface and solving the appropriate partial differential equations,
the measured surface information can be transformed to yield the load-displacement profiles at
any point along the rod-pump system. Gibbs and Neely showed the load-displacement diagram at
the two points along a tapered rod string where the rod sizes changed ( Figure 9).

Figure 9

The first such chart is at a depth of 1,475 feet. The transformation of the dynamometer card to
this depth shifts the position of the curve to lower loads and shorter displacements as
expected.
The second transformation shows a load-displacement at a depth of 3,125 feet.
The third transformation is made for the pump depth of 8,525 feet. This, in essence, is the
synthetic dynagraph.

In (Figure 10), the scale of the dynagraph has been expanded to show that the fluid load is 3,200
pounds. The buoyant force on the rods is given by the magnitude of the negative load on the
dynagraph. Here, it is 1,400 pounds.

Figure 10

The displacement dimensions show that the gross pump stroke is 7.1 feet, but that the net liquid
stroke is 4.6 feet. This means that the pumping efficiency, based on the relative displacements, is
approximately 65% (4.6 feet / 7.1 feet) and that some gas compression is occurring on the
downstroke. It also shows that the tubing anchor is working well to hold the tubing during the
upstroke. With knowledge of the pump diameter and the pump speed, the displacement at the
pump is calculated to be 200 B/D. The surface production rate was measured at 184 B/D and so
no serious tubing leaks are evident.
With the dynamometer cards calculated at the top of each tapered rod section the load on the top
rod of each section can be measured; dividing that load by the rod cross-sectional area gives an
estimate of the respective rod stress. Doing this yields a peak rod stress of about 30,000 psi, a
value that is below rod limitations. The dynamometer card is used for the balance of the analysis.
When synthetic dynagraphs are not used, an alternative is to use the dynamometer card to
interpret both the surface and subsurface operation of the pumping system. As a guide to this type
of interpretation API Bulletin 11L2 (1969; reaffirmed Sept. 1999) provides an extensive set of
typical dynamometer cards, categorized according to the dimensionless variables used in the
pump design procedures.

Progressive Cavity Pump (PCP) Systems


PCP SYSTEM OVERVIEW

A progressive cavity pump (PCP) system has three main components (Figure 1: PCP system.
Courtesy Weatherford International Ltd).

Figure 1

A surface drive unit, powered by a gas or electric prime mover, direct-drive electric motor or
hydraulic motor.

A rod string, which connects the surface drive unit to the subsurface pump.
A positive displacement subsurface pump.

A progressive cavity pump consists of a single helical rotor, which turns inside of a double-helical,
elastomer-lined stator (Figure 2: Rotor-stator assembly. Courtesy R&M Energy Systems ).

Figure 2

In a typical installation, the rotor is connected to the rod string, while the stator is run on the end of
the production tubing.* To operate the pump, the surface drive unit rotates the rod string, which in
turn moves the rotor. The eccentric rotation of the rotor within the stator causes a series of sealed
cavities to form. These cavities move progressively from the suction end to the discharge end of
the pump body. As each cavity diminishes in size, another increases in size at the same rate.
Thus, the total cross-sectional area of the cavities is constant. This results in a continuous, nonpulsating, positive displacement of fluid. The fluid flow rate is proportional to the size of the cavity,
the rotational speed of the rotor and the pressure differential from the inlet to the discharge end of
the pump.
* Exceptions to this "typical" PCP installation include

Insertable PC pump systems (e.g., R&M Energys IPCTM): The rotor


and stator are run on the rod string to fit inside the tubing as a
single unit. This system, which can be installed in tubing diameters
ranging from 2 7/8 inches to 5 1/2 inches, allows the pump to be
installed or retrieved without having to pull the tubing string.

Electrical submersible PC pump systems. Rather than driving the pump from the
surface, a downhole ESP motor is coupled with a gear reducer to rotate the
pump. This eliminates the need for a rod string, and works to an advantage in
deviated wells where rod where is a concern. The pump can be run and retrieved
using a slickline unit (Lea and Winkler, 1998).

HISTORY AND DEVELOPMENT


Ren Moineau designed the first progressive cavity pump in 1930 during his doctoral studies at
the University of Paris. In 1932, he and Robert Bienaim co-founded PCP Pompes, which became
the first commercial provider of these pumps and has been producing them ever since. Then, as
now, surface pumping was a major area of application. Because they have few moving parts and
provide a non-pulsating fluid displacement, PCPs are especially suitable for moving heavy,
viscous and/or solids-laden liquids.
The first subsurface application of Moineaus concept was the development of downhole drilling
motors or mud motors, which began in the 1950s. These motors actually operate on a reverse
Moineau principle. In other words, rather than using mechanical energy to turn the rotor and
displace fluid, drilling fluid is pumped downhole to turn the rotor and the drill bit to which it is

connected (Figure 3: Downhole mud motor). The ability to turn the drill bit without rotating the drill
string has proven invaluable in wellbore deviation control and directional drilling.

Figure 3

In the early 1980s, as operating companies were looking for ways to pump heavy oils in under
viscous, high-solids conditions, attention turned to progressive cavity pumping as a lift method. It
was then that Robbins & Myers, Inc. designed a PCP system for use in artificial lift (Reyard, 1995).
In this application, the stator is run on the end of the production tubing, while the rotor is attached
to a rod string, which in turn is connected to a wellhead drive assembly powered by a rotating
surface power source.
Subsequent innovations in PCP technology have led to its use in a growing range of applications
until, in many areas, it stands alongside beam pumping, electrical submersible pumping and
hydraulic pumping as a viable artificial lift option.

AREAS OF APPLICATION
PCP systems are particularly suited for producing high-solids and/or heavy crude oil and bitumen,
and for high water-cut wells, including those that produce from mature waterfloods. They may also
be used for dewatering coalbed methane and other types of gas wells. They can be used to
produce medium-gravity crude oils (subject to limits on H 2S and CO2 concentrations) and light
crude oils (subject to limits on their aromatic content).

Progressive cavity pumping offers a number of advantages relative to other artificial lift methods
(Saveth and Klein, 1989; Weatherford, 2003):

Low initial capital costs


Ease of installation
Portable and lightweight to transport
Low power consumption
Minimal maintenance requirements
High system efficiency
Good solids handling capability
Few moving parts
No internal valves to become sanded or gas-locked
Wide range of flexibility (production rates from 2 to 4500 Bbl/D; oil gravities from 5 to 40
degrees API; pump speed can be adjusted to match production rates by changing sheaves,
or by using hydraulic drives or variable frequency drives. Suitable for vertical, deviated or
horizontal wells)
Low surface profile (well suited for areas with height restrictions)
Low noise levels
Low tendency for fluid leakage at the wellhead.

Disadvantages of PCP systems include the following:

Limited depth range (typically around 2000 to 4500 ft, with a maximum of 6000 ft); due to the
torsional strength limits of the sucker rod string that turns the rotor.

Limited temperature range (typically 75 to 150 F, with a maximum of less than 275-300 F),
based on the material used in the stator elastomer.

Degradation of elastomer properties due to aromatic compounds in higher-gravity crude oils,


presence of H2S, or use of hot-oil or chemical paraffin treatments

The equipment used to manufacture the rotor and stator limits the maximum element length to
about 200 inches.

Not as familiar to field personnel as other artificial lift systems.


Lower utilization of PCP equipment affects salvage value.

PCP SYSTEM COMPONENTS


This section describes the major components of a PCP system, including the subsurface pump,
the rod string and the surface drive unit.

SUBSURFACE PUMP
The heart of a PCP system is the rotor-stator assembly that makes up the subsurface pump
(Figure 1: PCP rotor-stator assembly. Courtesy R&M Energy Systems).

Figure 1

The rotor is a precision-machined, single external threaded helical gear made of high-strength
steel and coated with hard chrome plating to protect against abrasion and minimize friction. For
extremely abrasive conditions (e.g., high-rate wells that produce more than 10 percent sand),
pump manufacturers may apply alternative rotor coatings using thermal spray processes (Mills
and Gaymard, 1996).
The stator is a tube with an internal form like a double helix. The internal surface of the stator is
coated with an elastomer that is bonded to the base metal. In highly corrosive environments, an
all-stainless steel rotor and stator might be used. The rotor and stator have the same minor
diameter, but the distance between the threads, the pitch length, of the stator is twice the pitch
length of the rotor.
As the rotor turns, the rotor and stator form a series of sealed cavities, 180 degrees apart, which
move from the suction end to the discharge end of the pump. As one cavity gets smaller, the
opposing cavity enlarges at the same rate; this keeps the fluid moving at the same pressure and
speed.

STATOR ELASTOMERS
An elastomer, as defined by the American Heritage Dictionary, is any of various polymers having
the elastic properties of natural rubber. Or, as noted in Mills and Gaymard (1996), it is a material
that can be stretched repeatedly to at least twice its original length and return rapidly to virtually
its original dimensions.
The elastomer that coats the internal surface of the pump stator is a critical PCP component. It
provides a seal between the rotor and stator that allows progressive cavities to form, and allows
abrasive particles to pass through the pump or become embedded rather than damage the stator.
The material selected for the elastomer is determined by well operating conditions, including the
API gravity and composition of the oil, the operating temperature and concentrations of water,
CO2, H2S, free gas and solids. Incompatibility between the elastomer and the well fluid is a
common cause of PCP failures (PTTC, 2000).

Elastomer Materials
The standard PCP elastomer used in artificial lift applications is nitrile elastomer, or NBR, a
copolymer of butadiene and acrylonitrile (ACN). Butadiene provides elasticity and low-temperature
flexibility to the compound, while ACN imparts hardness, tensile strength, resistance to abrasion
and resistance to degradation by polar fluids such as oil. Concentrations of ACN range from
around 20 percent to just over 50 percent. As the ACN content increases, resistance to swelling
increases, tensile strength, hardness, and abrasion resistance increases, and high-temperature

performance improves. However, the tradeoff that accompanies an increase in ACN content is the
decrease in elastomer resilience and permeability qualities (Revard, 1995).
Hydrogenated Nitrile Elastomer (HNBR) is nitrile in which the hydrogen saturation of the polymer
has been increased. This increases the elastomers resistance to chemical and temperature
degradation.
Non-nitrile compounds include fluorine-containing polymers, which have been developed for use
in high-temperatures and higher-gravity oils.

Elastomer Performance Characteristics


Pump manufacturers categorize stator elastomers in terms of their performance characteristics
and general areas of application (e.g., Tables 1 and 2). Specific applications are based on
evaluating individual operating conditions and on laboratory compatibility testing.
Table 1: Elastomer Performance Characteristics
(After R&M Energy Systems, 2003)
Type

ACN
Level

Resistance
Aromatic
well fluid

H2S

Gas
permeati
on

Tensile
Streng
th*

Tear
Resista
nce*

Max.
Tem
p.

NBR

Mediu
m
(2735%)

Acceptabl
e

Good

Superior

100

100

175
F

NBR

UltraHigh
(4251%)

Good

Barely
accepta
ble

Superior

95

92

175
F

HNBR

UltraHigh
(4251%)

Excellent

Excellen
t

Excellent

152

93

250
F

HNBR

High
(3641%)

Good to
excellent

Superior

Excellent

148

107

250
F

FKM

N/A
Flouro
Elasto
mer

Superior

Barely
accepta
ble

Not
recomme
nded

60

69

375
F

* Normalized to 100%

Table 2: Elastomer Performance Characteristics


(After PCP Pompes, 2006)
Resistance to
Type

Mechanica
l
Properties
*

Abrasive
s

Aromatic
s

H2S

CO2

Temp.
Limit

Standard
Nitrile

Excellent

Good

Medium

Good

Good

120
C

Soft Nitrile

Good

Excellent

Poor

Good

Poor

80 C

High ACN

Good

Medium

Good

Good

Medium

100
C

Hydrogenate
d Nitrile

Good

Excellent

Medium

Excellen
t

Medium

140
C

Fluorocarbon

Medium

Poor

Very
good

Good

Excellen
t

130
C

*Elasticity, tear resistance

ALL-METAL PCPS
An inherent shortcoming of progressive cavity pumping is the temperature limitation imposed by
the materials used in the stator elastomer. This precludes its use in reservoirs that produce under
thermal enhanced recovery. This is unfortunate, because PCPs are otherwise well suited for
pumping the types of heavy oils that thermal methods are designed to produce.
This limitation has led to the development of an all-metal progressive cavity pump as part of a joint
effort by PCM Pompes and TOTAL (Beauquin et al, 2005)as of 2005, this development was in
the advanced testing stages.

ROD STRING
Most PCP systems employ API sucker rods. But unlike a beam pumping system, where the rod
string is mainly designed to withstand reciprocating axial loads, the primary load on a PCP rod
string is torsional. Thus, rod torque is a primary design consideration in progressive cavity
pumping.
Depending on the pump depth and flow requirements, conventional sucker rods may be more than
adequate to drive the pump. If they are not, high-torque rods (e.g., Weatherford T-Rod, with
specified torque limits of 1600 lb-ft for a 1 x 1-inch rod, and 2500 lb-ft for a 1 x 1 1/8 inch rod)
may be incorporated into the design. Hollow rods with maximum operating torques ranging from
1000 to 2500 lb-ft are also available for such applications Olmos et. al, 2001)for example, the
PCPRod, manufactured by Tenaris (Figure 2. Courtesy of Tenaris Oilfield Services, 2003).

Figure 2

TUBING AND ROD STRING ACCESSORIES


Tubing and rod wear are major concerns in both reciprocating and progressive cavity pumping
systems, particularly in crooked or deviated wells where the rods and tubing are most likely to
come into contact. Some of the devices that are used to minimize rod and tubing wear, facilitate
PCP installation and maximize operating efficiency include:

Torque Anchors: Torque anchors are used where there is potential for high torque, elastomer
swell, heavy oil or a likelihood of backing off from the tubing. They are installed on the
tubing to reduce tubing/stator motion and resultant wear against the casing, and to prevent
backing off due to rotational torque. ( Stop Bushings: A stop bushing is installed on the
lower end of the stator so that the rotor can be properly spaced out during pump installation.
Gas Separators: A PC pump is usually set below well perforations so that the gas breaks out
above the pump intake. However, an alternative installation involves placing a perforated
tubing joint and dip tube below the pump intake, thus causing gas to break-out before the
fluid enters the pump. This will maintain high pump efficiency.
Tubing Rotators: Installed at the surface, these are used in both beam pumping and PCP
systems to periodically rotate the tubing string and prevent uneven wear. This reduces the
frequency of tubing failures, thereby decreasing downtime and workover costs. Tubing
rotators are available with electrical, mechanical and manual drive options (Figure 3 : Tubing
rotator. Courtesy R&M Energy Systems).

Figure 3

Rod Lifter: rod lifter is a portable hand tool installed over the polished rod extension. It is
designed to raise the rod string by some desired amount, utilizing the power from the PCP
surface drive unit. This allows for easy repositioning of rod couplings so as to more evenly
spread wear along the tubing string, and resetting of subsurface pumps that may be
improperly placed (Figure 4: RodecTM Rod lifter. Courtesy R&M Energy Systems).

Figure 4

Rod Guides: Rod guides of various types are available for use in both beam pumping and
PCP systems (Figure 5: Rod guides. Courtesy R&M Energy Systems). They are used to
centralize the rod string (thus preventing rod and tubing wear) and to prevent paraffin
buildup.

Figure 5

Anti-Friction Unit (AFU): An anti-friction unit is installed when excessive friction is anticipated
due to solids in the fluid (typically experienced in dewatering operations for coal-bed
methane wells). The AFU works by using two hydraulic cylinders mounted on the pump drive
assembly to lift the rod string and rotor when torque is detected. This unit has also been
installed in heavy oil or tar sand applications where pumps become sand-locked.

SURFACE DRIVE UNIT


The principle components of the surface drive unit are the drive head and the prime mover
(Figure 6: Surface drive unit for PCP system. Courtesy R&M Energy Systems ).

Figure 6

DRIVE HEAD
The drive head is installed directly on the wellhead. Its functions are to support the weight of the
rod string, transfer rotational energy from the prime mover to the rods and seal the surface

equipment from the produced well fluid. The drive head is also provided with mechanical or
hydraulic controls to regulate recoil or backspin of the rods.
Drive heads are sized in terms of maximum horsepower rating, thrust bearing capacity, polished
rod diameter, maximum polished rod speed and maximum polished rod torque. For example, the
operating parameters for three Moyno Ultra-Drive heads are as follows (R&M Energy Systems,
2003):
Model:

AA4

CV1

EX1*

15 hp

100 hp

300 hp

Dynamic thrust rating (Ca90),


standard thrust bearing:

13700 lb

33700 lb

80400 lb

Available polished rod diameters:

1.25 inch

1.25 and 1.5


inch

1.5 and 2.0625


inch

Maximum polished rod speed:

450 RPM

600 RPM

600 RPM

Maximum torque:

500 ft-lbs

1750 ft-lbs

4000 ft-lbs

Maximum power rating:

*EX1 can employ dual electric motors


The most common type of drivehead assembly is a hollow shaft design that uses a polished rod
and polished rod clamp. This design allows the polished rod to move through its bearing housing
and the stuffing box assembly. This type of drivehead assembly also makes it easier to space the
pump and adjust it without removing the drivehead from the wellhead (as must be done with a
solid shaft design). Repairs are also easier.
The drive head may work by direct drive, using a gear reducer or a belt/sheave assembly, or it
may employ a mechanical, electronic or hydraulic variable speed drive.

PRIME MOVERS
Electric motors may be used to power any of the surface drive assemblies designed for PCP
systems, and they offer a number of advantages. They are clean, quiet, efficient, smooth-running
and require very little maintenance. They are well-suited for automatic operation, and where flow
rates are likely to change, they can be used with variable frequency drives.
Natural gas-powered engines can also be used as prime movers. These require more equipment
than electric motors (i.e., right-angle shaft, belts and sheaves), cannot be used with variable
speed drives or torque monitoring and shutdown systems and may encounter vibration problems.
However, they may be the option of choice where gas supplies are plentiful and electricity supplies
are limited.

FLOW AND PUMP-OFF CONTROLS


Flow control and pump-off control devices are used to prevent the PC pump from running dry and
burning out when low or variable production rates are expected, or when fluid levels must be kept
at or near the pump intake
Flow controls typically receive input from:

Flow line pressure sensors


Production line flow sensors that detect volume or temperature changes
Torque sensors
Time clocks

PCP SYSTEM DESIGN AND INSTALLATION


Like other positive displacement pumps, the selection and sizing of a progressive cavity pump are
based on its displacement and lift capacity.

The displacement of a PCP is the product of its cross-sectional area and the fluid velocity,
which may be expressed as (R&M Energy Systems, 1997):

q = K x 4E x Drotor x Ps x N
Where q = flow rate
K = conversion factor
E = eccentricity of rotor
Drotor=rotor diameter
Ps=stator pitch
N=pump speed (RPM)

(1)

The lift capacity or pressure capability of a PCP is a function of the number of pump stages, or
the number of times the progressing seal lines are repeated, and is affected by the viscosity
of the fluid being pumped and the compression fit between the rotor and stator (Saveth and
Klein, 1989).

Manufacturers specify their pumps capabilities based on these two parameters. For example, a
Moyno Model 50-H-400 PCP is rated for a displacement of 4 BFPD /RPM, and a lifting capacity
of 5000 feet of water (R&M Energy, 2002).

GENERAL DESIGN CONSIDERATIONS


If an initial screening indicates that progressive cavity pumping is a viable artificial lift option, a well
data sheet should be prepared to help decide what model of pump is suitable for the expected
operating conditions. This sheet may be used to provide input data for the vendors software or
other commercially available design package. Some of the key input parameters in this evaluation
process include:

Required fluid flow rate and total dynamic head: Manufacturers can design PCP systems for a

wide range of flow rates and lift capacities, based on a production system analysis of the
proposed installation.
Depth of the producing zone: This will affect the pump specifications, the design of the rod
string and the power requirements for the surface drive unit.
GOR and water cut: Published lifting capacities are usually referenced to water. NBR and
HNBR elastomers exhibit superior resistance to gas permeation.
Abrasive applications: With a PCP, solid particles tend to either pass through the pump
cavities or become embedded in the flexible elastomer, rather than plugging or causing
abrasion to the pump elements. This makes progressive cavity pumping the lift method of
choice for high-solids, abrasive fluids. To further minimize abrasive wear, a pump design
may incorporate (1) special rotor coatings applied using thermal spray processes, (2)
selection of elastomers that are soft and flexible enough to allow particles to pass through
seal lines without damaging the stator. The pump should be designed to produce at the
desired rate while maintaining the lowest fluid internal velocity and the lowest possible
rotating speed (R&M Energy Systems, 1997).
Fluid composition, including aromatic content and presence of H2S or CO2: This will affect the
choice of elastomer.

ROLE OF PUMP SUPPLIERS

PC pump manufacturers typically market their pumps directly to producers. A major barrier to the
expansion of the PC pump market has been the lack of knowledge and experience in the proper
use and maintenance of these pumps, resulting in high failure rates. Be sure to choose a pump
supplier with a proven record of ability and service.
The pump supplier should also make pump performance tests and match the rotors and stators to
the well conditions. A competent pump supplier will also perform power consumption studies so
that economic performance of the well can be monitored.

PUMP DEPTH
The most important factors that determine the setting depth of a progressive cavity pump are:

Dynamic fluid level


Presence of solids in the fluid
Depth of production perforations relative to well total depth

PC pumps are usually set below the perforations in oil and gas wells, especially when dewatering
coalbed methane gas wells. This maintains high volumetric efficiency by reducing the amount of
gas moving through the pump.
In wells with a strong water drive or in waterflood wells, the pump can be set higher in the wellbore
if the fluid level is high and the risk of running the pump dry is low. The pump also may be set
above the perforations if the well has no rat hole or if the fluids contain solids.

PCP INSTALLATION AND STARTUP


Installing a PC pump begins with the stator. The stator is attached to the bottom tubing joint. The
stator is usually threaded on both ends to allow gas anchors or tubing anchors to be attached. The
tubing is typically run deep enough in the hole to place the stator below the perforations.
The rotor is run into the well on a full-length sucker rod at the end of the rod string. The weight of
the rod string should be recorded just before moving the rotor into the stator. When the rotor
moves into the stator, the rod string may rotate to the right and the rod weight will decrease.
Once the rotor is in place, the rod string is picked up so that the rods are in tension. This is the
operating point for the pump. The next step is to account for the length of the shaft of the
drivehead, pull enough sucker rods to make room for the polish rod and finally, install the
drivehead onto the polish rod.
The most common type of drivehead assembly is a hollow shaft design that uses a polished rod
and polished rod clamp. This design allows the polished rod to move through its bearing housing
and the stuffing box assembly. An important safety precaution is to install a drivehead with a
braking device. Sometimes a PC pump will rotate backward when shut off, and a braking device
safely controls that "backspin".
This type of drivehead assembly also makes it easier to space the pump and adjust it without
removing the drivehead from the wellhead (as must be done with a solid shaft design). Repairs
are also easier.
It is important to follow a pre-start checklist when installing a PC pump system. This checklist
should cover at least the following steps:

First, keeping the service rig on location until the pump is working properly. If the pump must

be reset, the service rig will be needed to re-space it.


Measuring and recording the fluid level in the well before starting the pump.
Checking the prime mover rotation to make sure that the rod string can rotate clockwise.
Making sure that any belts and sheaves are aligned and that the belt guard is in place.
Making sure that bearings, brake, and packing are properly lubricated. Don't tighten down the
packing until the well has pumped up.

Making sure that all flow line valves are open before starting the pump.
Measuring and recording the fluid level in the well after startup and comparing it with the level
before startup.

PCP OPERATION
Pump performance testing is an important aspect of production optimization. Since the key to
successful PC pump operation involves matching the pump performance to the well conditions, it
is necessary to test a new pump to determine the best rotor and stator combination. The pump
also should be re-tested whenever it is pulled, to check for wear and to ensure that it is operating
at its highest possible efficiency.

INITIAL PERFORMANCE TESTING


Rotors and stators should be matched to the intended well conditions. When installing a new
pump, it is important that the rotor and stator set have been tested together as a set (rather than
using test-stand tests of the downhole equipment), and that the expected fluid temperatures and
operating pressures are considered. Such testing ensures the best match of the actual equipment
to be used and reduces the chances that parts will not fit and operate properly. The goal of the
initial performance test is to achieve long pump run times while reducing lifting costs.
A good initial pump performance test includes evaluations of

Mechanical and volumetric efficiencies: Pump efficiency changes with operating pressure and
temperature.

Measured and theoretical torque: The difference between the measured and theoretical
torque shows actual internal friction, which affects mechanical efficiency and power
requirements
Production and pressure at a given RPM: Compare actual with design performance.

For easy future comparisons, these results should be displayed in a curve format.

PUMP INSPECTION AND RETESTING


After the rotor and stator are pulled out of the well, the pump diameters should be measured and
the pump elements visually inspected.

ROTOR INSPECTION
Abrasive wear is typically found along the contour of the rotor, whereas normal wear is more
evenly distributed over the entire rotor. The rotor diameter should be measured and the surface
should be examined for signs of corrosion. Look at the top, bottom and sides of the rotor for any
marks, discolorations, burred metal, or burned rubber.

STATOR INSPECTION
The stator should be visually inspected for signs of swelling or dry run by looking at the internal
diameter and determining if the elastomer is brittle or cracked, or if there are chunks of elastomer.

RETESTING
Whenever a pump is pulled after a long run time or after a problem has been reported, it should be
retested to compare its current performance with its original performance. If one of the pump's
elements (the rotor or stator) fails and must be replaced, then test the used element with the new
replacement part to see if the pump performance is still satisfactory for the well conditions.

If the elements work well together, the expense of retesting will have been warranted by insuring
that the reinstalled pump will function as expected. If they dont work together, the cost of installing
and then re-pulling the pump will have been eliminated.

SYSTEM MONITORING
The performance of a PC pump should be checked regularly. The pump should be tested to
monitor volumetric efficiency, dynamic fluid level, pump speed, flowline and casing pressure, and
electric motor power use.

MONITORING PUMP PERFORMANCE


Performance and production data from a pump in service can be compared to the data for the new
pump. Little maintenance is required during operation, other than checking well productivity.
To create a comparison baseline, pump performance should be tested before installation and at
any time thereafter when the pump is pulled out of the well.

The rotor should be inspected for wear from normal operation, abrasion and corrosion
The bottom of the rotor should be checked to ensure that it is not riding on the stop pin
The stator should be visually inspected:
o Check the stator elastomer surface to see if it is affected by operations or well
fluids
o Gauge the internal surfaces to check for wear or swelling of the elastomer

DRIVE ASSEMBLY MAINTENANCE


Drive assemblies are made up of a radial bearing and a thrust bearing. Some drives include a
right-angle shaft with a reduction gear. Most drives also have a backspin brake. The drive
assembly includes a stuffing box to seal around the polish rod in order to prevent fluid leaks.
Normal maintenance of drive assemblies consists of changing the oil, greasing the stuffing box,
and checking the tension on any belts checking the motor mounts and on belt guards.

REMOTE MONITORING OF PCP WELLS


Remote monitoring and control techniques are often applied to wells that use PCPs. These
systems can measure the performance of a single well or an entire field using sensors and data
transmitting devices that tell field personnel when pump performance has changed. By comparing
measurements of pump performance over time, the operator can analyze and optimize the
operation of the PCP and the well itself, thus improving efficiency, and better managing field
operations.
One feature of modern monitoring systems is the ability to remotely measure pump performance
and change pump activity using two-way control devices. By continually monitoring and comparing
parameters like electricity usage, surface drive unit speed, rod string torque, fluid movement,
pump intake and discharge pressure, and bottom hole pressure, the operator can remotely identify
potential problems and then take preventive or corrective action. Through monitoring, problems
such as paraffin buildup, perforation scaling, sanding, pump starvation, formation skin damage,
and surface flow restrictions can be identified. Many systems are capable of automatically
notifying the operator when key parameters vary beyond specified limits.

Monitoring System Components


Key components of the monitoring system include:

Downhole Pressure & Temperature Sensors


Sensor Data Process System

Well Controller (VFD)


Remote Terminal Unit (RTU)

The permanent sensors measure intake and discharge pressure and temperature so that reservoir
conditions can be evaluated. A sensor data process system allows the sensors to communicate
with well controllers, (i.e., a vector flux drive, which provides speed and torque control). This lets
the operator control the well based on changing downhole conditions.
A remote terminal unit (RTU) is used to transmit data continuously or store well performance data
for later transmission and analysis. The RTU also allows the operator to communicate back to the
well.
PC pump monitoring systems have paid off in increases in reservoir, rod string and pump life.
Furthermore, the continuous data acquired by such systems is often used to design future PCP
installations and reduce recurring pump problems in a well.

PCP TROUBLESHOOTING
When a problem occurs with a PCP installation, the first step in troubleshooting is to determine
how the operating conditions may have changed. A good place to begin is to review the records of
pump operations such as electrical power usage (amps) and fluid levels recorded during
installation. Comparing the initial data with the current data may help identify the problem.
The next step is to make sure that the surface equipment is operating properly. This involves
checking the drivehead, prime mover, electric motor, power lines (low voltage), fuses, hydraulic
drives, and variable frequency drives.
If a PC pump fails because of problems with the rotor or stator rather than the sucker rod string, it
may be necessary to replace those components. When a pump wears to the point that the fit of
the rotor and stator becomes too loose to lift the fluid column, the stator must be replaced. Usually,
the stator wears out before the rotor because the stator elastomer wears out before the rotor does.
A rotor can sometimes be re-chromed and re-used, but a stator cannot be repaired, and must be
replaced.

RUN DRY OPERATION


This is a common mode of failure for a PC pump, which occurs when the pump runs dry after
operating in a pumped-off condition for an extended time. When there is not enough fluid for
lubrication, friction inside the pump increases, causing the elastomer surface to scorch or burn.
This causes the pump to lose its ability to seal and then prevents the pump from lifting fluid to the
surface. When this happens, the stator elastomer surface will look hard, cracked, and crumbly and the internal diameter of the stator will appear larger than normal. The rotor also may look
burned or have elastomer on it.
One way to avoid running the pump dry is to adjust the pump operating speed. When installing a
new pump, set the pump RPMs at a lower level and monitor the fluid level. Gradually increase the
pump speed until it delivers the desired production rate. Another way to avoid running the pump
dry is to use a pump off controller, which will shut off or slow down the pump.

GAS PERMEATION
PC pumps can fail if gas saturates the stator elastomer, making the elastomer surface blister. This
increases pump torque and internal friction and may cause the sucker rods to part. It is also
possible for the torque to break the rotor. When gas permeates the stator elastomer, blisters or
bubbles will be visible on the surface of the elastomer. The bubbles can range from very small to
very large. If the blisters or bubbles are large, the elastomer may break away. Sometimes, the
blistering may not appear until the stator has been out of the well for some time.
If gas is present in the well, the preferred installation is to set the pump below the perforations so
that the gas can break out before entering the pump. If the pump can't be set below the

perforations, a gas anchor can be used to reduce the amount of gas entering the pump. If gas
permeation is anticipated, its effects may be reduced or eliminated by using a different stator
elastomer.

SWELL
A stator can swell as a result of reactions with aromatics in the produced fluids or chemical
treatments used in the well; or because of high bottom hole temperature or gas permeation.
Swelling increases the pump torque and may cause parting of the sucker rods, broken rotors, or
electrical overloads. High torque can even cause the tubing to come unscrewed.
Before a pump fails, swelling might be noticed by monitoring the amperage, which is a good
indicator of torque. After pulling the pump, it is good to conduct a performance test to measure
pump torque and to visually inspected the pump.

HYSTERESIS
This type of failure occurs when the elastomer is overpressured from too much internal friction.
This makes the elastomer mechanically break down and tear away from the stator. This condition
may occur when a pump is run dry, or when the rotor is not engaged fully, resulting in
overpressuring of the part of the pump that is engaged.
Running a pump dry often initiates a hysteresis failure. Since a dry run begins at the pump intake
and moves upward to the discharge end, the upper section becomes overpressured and fails.
Visual inspection of the stator will show chunks of rubber that have torn from the tube and
evidence of tearing of the elastomer.
To prevent hysteresis failure, make sure that the pump is correct for the application and that there
is no overpressuring in the downhole or surface system.

BOND FAILURE
This type of failure is the result of a true manufacturing defect. Bond failures occur if the elastomer
bonding agent is weak or is improperly applied to the stator tube before the elastomer is molded.
When the bond fails, the stator elastomer will detach from the stator tube, again causing high
torque and catastrophic failure. A bond failure is indicated when the elastomer comes out in large
sections rather than in chunks as with a hysteresis failure.
Visual inspection of the stator tube will show that the interior of the tube has no adhesive on it. The
clean area may cover only a small area.

ROTOR FAILURES
PC pump rotors fail for a variety of reasons. Some of the most common failure types are described
below:

Abrasion and Solids


If a PC pump is operating at high RPMs in a well containing abrasive fluids, the rotor may fail due
to excessive wear. This failure can also occur if solids are present in the well fluids. When the rotor
is worn from abrasion, the pump loses its internal fit and cannot lift the fluid column.
A pump that has produced abrasive fluids will show wear along the length of the rotor. To avoid
this kind of wear, it is best to install an oversized pump and run it at lower RPMs to move the
solids without damaging the rotor.

Rotor Wear
When the rotor is pulled, it is not uncommon for the chrome to be missing from the top 6 to 12
inches on the rotor profile. This is usually due to friction between the rotor and the tubing
immediately above the stator. This condition does not create an operating problem since this part

of the rotor is not engaged into the stator; this condition may also indicate that the tubing is the
wrong size and may increase torque and vibration.

Pitting and Corrosion


If the pump is installed where corrosive fluids are present, the chrome plating and the base metal
of the rotor can become pitted. This will cause wear on the stator elastomer. Usually, corrosionpitted rotors cannot be re-chromed, and must be replaced.

Chrome Flaking
Flaking is another result of corrosion, but unlike pitting, the chrome flakes off. Corrosion forms
beneath the chrome, causing flakes of chrome to come off. In some cases, flaking may be a
manufacturing defect. In cases of pitting and flaking, it might be better to replace the chrome rotor
with a stainless steel rotor.

Rotor Breaking
In almost all cases, rotors break because of high torque; but in some cases, manufacturing
defects may also be responsible. In addition of the previously mentioned causes for high torque
swell, blistering and bubbles, and the presence of solidsthe pump design may be responsible for
high torque if the rotor and stator are mismatched. This problem is noticeable when the pump runs
at low RPMs, creating excessive vibration.

Hydraulic Pump Systems


HYDRAULIC PUMPING: OVERVIEW
A hydraulic pumping system is made up of the following components:

Surface prime mover


Surface pump
Power fluid system and control manifold
Subsurface pump, which is driven by the power fluid.

The prime mover provides power to the surface pump, which pressurizes and injects power fluid
downhole through an isolated tubing string. This power fluid actuates the downhole pump and
causes it to move fluid to the surface.
The power fluid system may be either closed or open. In a closed system, the power fluid is not
mixed with the produced fluids, and a separate downhole power fluid line is needed to return the
spent power fluid to the surface (Figure 1: Hydraulic pumpingclosed power fluid system).

Figure 1

In an open system, the power fluid mixes with the produced formation fluids during the pumping
cycle (Figure 2: Hydraulic pumpingopen power fluid system).

Figure 2

The power fluid may be water or oil. In either case, facilities for cleaning the power fluid must be
included in surface facility system designs.

SUBSURFACE HYDRAULIC PUMPS


There are two main types of subsurface pumps used in hydraulic pumping systems: the
reciprocating, or piston pump, and the jet pump. Although both types of pumps get their lift energy
from the power fluid that is injected from the surface, they operate on very different principles.

RECIPROCATING HYDRAULIC PUMPS


In its simplest form, a reciprocating hydraulic pump is a closely coupled engine-and-pump
assembly (Figure 1: Reciprocating hydraulic pump). The unit is installed below the working fluid
level. High-pressure power fluid flows from the surface to the hydraulic pump through a conduit,
and the spent power fluid and production return to the surface through one or more additional
conduits. The high-pressure power fluid causes the engine-end of the pump to reciprocate, driving
the pump-end, which moves the power fluid and produced fluids to the surface.

Figure 1

OPERATING PRINCIPLES
There are many different types of hydraulic pumps, each with a unique design and operation. One
of the simpler designs is the Weatherford Kobe Type A pump represented schematically in
(Figure 2), with the engine-end of the pump shown above and the pump-end shown below.

Figure 2

Although the pump works as a single unit, its operation is easier to understand if we consider the
engine end and the pump end separately.

Engine End
Consider the engine-end of the pump at the start of its downstroke, when the piston is at the top of
the cylinder (Figure 3: Engine end at beginning of downstroke).

Figure 3
At this point in the pump cycle, fresh power fluid is just beginning to enter the cylinder above the
piston, while spent power fluid is beginning to be exhausted from the lower part of the cylinder.
Note the rod attached to the piston. This rod extends up into the power fluid and down to the
pump-end. The upper end of the rod controls the power fluid direction of flow; the lower end of the
rod causes the pump-end to move fluids to the surface.
The piston next moves downward within the engine cylinder (Figure 4: Engine end during
downstroke).

Figure 4
Note that the rod above the piston (the valve rod) has two smaller-diameter sections. One is just
above the piston and the other is at the upper end of the rod. These two sections provide control
for both the movement of power fluid and the stroke length of the pump. The stroke length
corresponds to the distance between these reduced-diameter sections on the valve rod.
At the end of its downstroke, the piston reaches the bottom of its travel within the engine cylinder
(Figure 5: Engine end at completion of downstroke).

Figure 5
At this point in the cycle, the upper reduced-diameter section of the valve rod allows power fluid to
be ported below the engine valve. The engine valve now has the same pressure applied to its top
and bottom surfaces. However, because the bottom of the valve has the larger surface area, a
larger force is applied to the bottom of the valve, causing it to move upward. In doing so, it opens
and closes ports, thereby reversing the flow direction of the power fluid within the engine, and
causing the piston to reverse direction and begin its upstroke ( Figure 6: Engine end at beginning
of upstroke ).

Figure 6

As the piston moves upward in the cylinder, the reduced-diameter section of the valve rod located
just above the piston is about to enter the engine housing ( Figure 7: Engine end during upstroke ).

Figure 7

When the piston reaches the end of its upstroke, the reduced-diameter section of the valve rod
provides a fluid connection from a point beneath the engine valve to a lateral port within the
engine housing. This lateral port is connected to the low-pressure side of the engine ( Figure 8:
Engine end at completion of upstroke).

Figure 8

Because the pressure of the power fluid below the engine valve is now lower than at the upper
end of this valve, the difference in pressures forces the valve downward and reverses the flow of
the power fluid. With the valve in its new position, the cycle starts and the piston begins its
downstroke. As the piston in the engine-end reciprocates, the connecting rod causes the piston in
the pump-end to move. This rod movement causes produced fluids to be pumped to the surface.

Pump End
The pump-end has a set of intake and exhaust valve balls at each end of the cylinder ( Figure 9:
Pump end of reciprocating hydraulic pump ).

Figure 9

On the downstroke, formation fluids enter the cylinder above the piston through the upper intake
valve. The fluids are then pumped out of the cylinder beneath the piston through the lower exhaust
valve.
On the upstroke, the valve positions are reversed. Produced fluids now flow into the lower part of
the cylinder and are exhausted from the upper part of the cylinder. Pumping occurs during both
halves of the pumping cycle. The opening and closing of valves and the pumping of produced
fluids continues with the engine-end and pump-end of the pump acting together.

ADVANTAGES AND DISADVANTAGES OF RECIPROCATING


HYDRAULIC PUMPS
Hydraulic piston pumping has a number of advantages relative to other lift methods:

Hydraulic pumps can produce at different fluid rates, typically from 135 to 15,000 B/D. The
pump speed can be easily changed to accommodate changing inflow conditions.

Pumps can be used over a wide range of depths, ranging from 1000 feet to 18,000 feet.
Free pump installations enable the pump to be circulated to bottom, and circulated back to
surface for repair or replacement, thus avoiding the need to pull the tubing.

Hydraulic piston pumps can be used effectively in crooked or deviated wells.


Wellheads and surface facilities are compact, thus making hydraulic pumping systems ideal
for offshore or urban production sites.

Chemicals to protect casing and tubing may be added to the power fluid, particularly in open
power fluid systems.

By restricting the amount free gas present, the hydraulic pump system can be among the
most efficient of all lift systems.

Disadvantages of hydraulic piston pumps include:

Potential fire hazard if oil is used for the power oil fluid
DIfficulty in pumping fluids with high percentages of solids
Negative effects on profitability if an oil inventory is required for a power oil system
Potential difficulty in diagnosing production problems when gas is not vented and pump
efficiency isSome installations require two strings of tubing
Dificulty in treating scale below the packer

HYDRAULIC JET PUMPS


A hydraulic jet pump has four main components: the pump body, the nozzle, the throat and the
diffuser ((Figure 10: Hydraulic jet pump).

Figure 10

Note that, like a reciprocating hydraulic pump, the jet pump is operated by the power fluid injected
from the surface. The difference is that it has no moving parts. Rather, it works by means of the

Venturi principle, transferring the energy of the injected power fluid to the produced formation fluid,
and then lifting both the power fluid and formation fluid to surface as follows (Petrie, 1987):
1. High-pressure power fluid is injected from the surface into the body of the pump.
2. The power fluid passes into the narrow nozzle. There, its potential energy is converted into
kinetic energyin other words, its pressure decreases and it becomes a high-velocity jet or
stream of fluid.
3. When the power fluid pressure drops below the pumps suction pressure, well fluid is drawn
into the pump, where it surrounds the power fluid jet.
4. The power fluid and well fluid are pushed into the throat, a straight cylindrical bore that has a
larger diameter than the nozzle exit. The power fluid and well fluid mix together in the throat,
and momentum is transferred from the power fluid to the well fluid. The mixed fluid stream
exits the throat, still at high velocity.
5. The fluid stream enters a diffuser. The expanding area of the diffuser results in a fluid velocity
reduction and a pressure increase. This increased pressure is used to lift the fluid to surface.

In the past, two companies manufactured jet pumps: Kobe and National Production. These
companies are now subsidiaries of Weatherford International Ltd. The operating principles are
similar, with the primary difference being the manner in which the fluid moves through the working
sections. In Kobe pumps, the nozzle points downward and receives power fluid directly from the
tubing string above it. The National Production Oilmaster pump has a nozzle that points upward,
and suction passages that do not reverse direction.

ADVANTAGES AND DISADVANTAGES OF JET PUMPS


Hydraulic jet pumps have a number of advantages, many of which are similar to those offered by
reciprocating pumps:

Jet pumps can be used in a variety of well conditions at fluid rates of 50 to 15,000 B/D and at
depths of up to 15,000 feet

Production rates can be increased or decreased by increasing or decreasing the power-fluid


supply

Because they have no moving parts, jet pumps can operate with low-quality power fluids and
can produce abrasive and corrosive well fluids

The jet pump can be hydraulically circulated to the surface for maintenance, thus reducing
downtime and expense

The jet pump is small, and can be run in many downhole settings, including deviated wells
Jet pumps move fluids at higher flow rates than reciprocating hydraulic pumps in the same
size tubing

Free-gas is handled better by jet pumps than by conventional pumps or centrifugal pumps
Multiple wells can be operated from the same surface injection facility

The main disadvantages of jet pumps are that they are usually less efficient than positive
displacement pumps, due to the high turbulence and friction that are created, and they are prone
to cavitation at low pump intake pressures.

DOWNHOLE INSTALLATIONS
RECIPROCATING PUMPS
The two basic types of hydraulic piston pump installations are fixed pumps and free pumps. Fixed
pumps are connected to the power fluid tubing and lowered into the hole; the power fluid tubing
must be pulled in order to retrieve the pump.

Free pumps are circulated into the well by pumping power fluid down the normal power fluid
conduit. The unit is pumped into the well with the standing valve closed. It seats on the bottom,
with the standing valve held open by a magnet, ready for operation. It is retrieved from the well,
when required, by reverse circulating power fluid down the return line, causing the pump to unseat
and travel to the surface.
Beyond these general classifications, hydraulic pumps can be installed in any of several downhole
configurations, and adapted to almost different tubing arrangements. Pumps are available for
production tubing diameters ranging from 1 1/4 inch to 4 1/2 inches.
There are four types of downhole tubing arrangements, and each of these may be designed as
open power fluid (OPF) or closed power fluid (CPF) systems.

FIXED INSERT PUMP


The tubing arrangement for a fixed insert pump with an open power fluid system is shown in
(Figure 1).

Figure 1

Note that the power fluid enters the pump through the power fluid line and that the spent power
fluid and production are commingled and produced up the annulus between the power fluid line
and the tubing. Gas is vented up the casing-tubing annulus.

FIXED CASING PUMP

With a fixed casing pump, the spent power fluid and production fluid travel upward between the
power fluid line and the casing (Figure 2).

Figure 2

In this arrangement, any formation gas must be handled by the pump, or a separate tubing string
must be installed to vent the gas from beneath the packer. Venting is normally required for high
GOR wells producing below the bubble point (Figure 3). Large pumps (for example, 3 13/16-inch
O.D.) are normally installed in this configuration.

Figure 3

PARALLEL FREE PUMP


(Figure 4) illustrates a parallel free pump arrangement with a closed power fluid system.

Figure 4

There are three parallel conduits within the pump: one for delivering power fluid, one to deliver the
spent power fluid back to the surface, and one to deliver the production to the surface. Gas is
vented through the casing. Because this is a free pump, the pump is unseated for retrieval by
reverse-circulating raw production. This causes a limited amount of produced fluids to contaminate
the power fluid system, but normally not enough to cause serious operating problems. Multiple
strings can be installed by running them simultaneously clamped together using a vertical set
packer or running strings separately in the casing-tubing annulus and stabbing into a suitable
connecting bowl socket. The size of free pump that may be installed depends on the casing and
tubing diameters. Table 1 lists some typical downhole arrangements.
Table 1: Free Pump Size for Various Casing and Tubing Diameters
(after Coberly, 1961)

Open System:

2-inch pump

2.5 inch pump

3 inch pump

4 inch pump

Tubing combination

2-inch EU
1-inch LP

2.5 inch EU
1.25 inch LP

3 inch EU
1.5 inch LP

4 inch EU
2 inch LP

4.51

5.58

6.79

8.375

Diameter of bundle, in.

Minimum casing size

5.5 inch, 20 lb/ft 5.625 inch, 24 lb/ft 7.625 inch, 24-lb/ft 9.625 inch, 47 bl/ft

Closed system:

2-inch pump

2.5 inch pump

3 inch pump

4 inch pump

Tubing combination

2-inch EU
3/4-inch LP
3/4-inch LP

2.5 inch EU
1 inch LP
1 inch LP

3 inch EU
1.25 inch LP
1.25 inch LP

4 inch EU
1.5 inch LP
1.5 inch LP

546

6.87

8.000

6.625 inch, 28
lb/ft

8.625 inch, 49 lb/ft

Diameter of bundle, in. 4.437

53.5 lb/ft
Minimum casing size

5.5 inch,
23 lb/ft

CASING FREE PUMP


(Figure 5) shows a casing free pump with an open power fluid system.

Figure 5

The commingled power fluid and production are produced up the casing-tubing annulus. Because
of its low cost, this is the most common type of hydraulic pump installation. Note that all free gas
must be handled by the pump unless a special vent line is installed to vent the gas from below the
packer. The pump is retrieved by reverse-circulating the commingled production and spent power
fluid. This will contaminate some of the fresh power fluid but, because it is soon commingled with
production anyway, it does not normally create an operating problem.

OTHER TYPES OF INSTALLATIONS


The fixed insert, fixed casing, parallel free pump and casing free pump are the four most common
hydraulic pump arrangements. But we could also use two pumps operating in tandem to double
the capacity of the equipment while using a single power fluid source. Alternatively, it is possible to
have tandem engines with a single pump, and tandem engines with tandem pumps.
It is also possible to hydraulically pump both zones of a dually completed well. In such a case, two
separate power fluid tubes are normally used because each completion will probably require
different operating-speeds and power fluid surface pressures. Another possible configuration is to
install a downhole safety valve, if one is needed, between the packer and the pump ( Figure 6)
Hydraulic pump installation with downhole safety valve). The actuating pressure for the safety
valve comes from the high-pressure line of the power fluid. If a critical event occurs at the surface,
the power fluid pressure will drop and the safety valve will close, preventing any further production.

Figure 6

JET PUMPS
Jet pumps can be adapted to fit interchangeably into the same downhole configurations as
reciprocating pumps (e.g., fixed insert and fixed casing), although some pump types require
special bottomhole assemblies. Like their reciprocating counterparts, they may be run as free or
fixed pumps depending on well conditions and production requirements.

SURFACE FACILITIES
The main surface components of a hydraulic pumping system are the power fluid system, the
pump system and the control manifold and wellhead system.

POWER FLUID SYSTEM


In a closed power fluid system, where the power fluid does not mix with the produced fluid , spent
power fluid returning from one or more wells enters the power fluid tank near its base. The power
fluid going to the wells is drawn off from near the top of the tank. A tank inlet for power fluid makeupvery likely coming from the surface processing facilitiesis also part of the system ( Figure 1:
Closed power fluid system).

Figure 1

Over time, the power fluid in a closed system may become contaminated with solid particles and
products of corrosion. For this reason, the returning power fluid in an oil system must be allowed
to settle in a power fluid settling-tank. When oil is used as the power fluid, an effective settling
system should provide an upward fluid velocity of less than one foot per hour, and the capability of
removing settled materials from the tank bottoms. When water is used as the power fluid, filters
may be used in place of a settling tank.
In an open system, oil enters the power fluid system from the production separator or treater, and
flows first into a gas boot to remove the last traces of gas from the treated oil ( Figure 2: Open
power fluid system) .

Figure 2

Gas flows from the top of the boot to the gas collection system, while dead oil flows from the
bottom of the boot into the power oil-settling tank. Oil going to the stock tanks is drawn off at the
vertical mid-point of the settling tank, and power oil is drawn off near the upper part of the settling
tank. Thus, most of the settling takes place between the point where production is drawn off and
the point where power oil is drawn off. Some lighter solids, then, may be drawn off with the
production, and heavier particles settle to the bottom of the tank, where they must be removed
from time to time. To be effective, the power oil tank should be large enough to allow the upward
velocity in the top half of the tank to be less than two feet per hour. This velocity should be lower

for heavier oils. If water is used as the power fluid, chemicals must be added for pump lubrication,
corrosion inhibition, and oxygen scavenging.
Jet pumps, because they work by mixing the power fluid and the well fluid, always employ an open
power fluid system.
A single well installation is used to accommodate an isolated well or for temporary operations. The
simplest form of this system uses a small surge tank with sufficient volume to allow settling,
enough pressure to charge the surface pump (20-100 psi), and a make-up pump to maintain a
fluid level. If the well has a free pump, then the tank should contain 2-3 times the tubing volume to
permit retrieving and running the pump (Figure 3: Surface hydraulic system for single-well
installation).

Figure 3

PUMP SYSTEM
Most surface pumps are skid-mounted and have either an electric motor or gas engine as the
prime mover (Figure 4: Kobe Triplex pump. Courtesy Weatherford International Ltd).

Figure 4

The Kobe Triplex pump shown here is one popular model, but any pump that can handle the lift
systems requirements may be used. Horsepower ratings of surface pumps normally range from
12 to 180 hp.
In addition to the pump elements themselves, the surface pump assembly includes a relief valve,
pressure gauges and safety switches.

CONTROL MANIFOLD AND WELLHEAD


The control manifold is normally supplied by the pump manufacturer. These multi-well distribution
manifolds are typically made in modular header sections that can be combined if new wells are
added to the injection system (Figure 5: Multi-well manifold).

Figure 5

This type of manifold usually contains pilot-operated control valves that maintain a constant
volume of power fluid going to each well regardless of pressure changes in the system. A
backpressure regulator is also used to maintain a constant pressure at the surface. Meters and
pressure gauges for each well are usually installed at this point. It is the control manifold that
normally controls downhole pump speed, performs pressure and engine-end efficiency checks,
and other troubleshooting activities.
The wellhead for a fixed pump must be designed to accommodate the power fluid and return lines.
The wellhead for a free pump is more complex: it must allow for the movement of power fluid
during pump installation and retrieval. This wellhead must also provide a means of catching and
holding the pump and safety device to prevent high-pressure from being applied to the casing. A
four-way valve at the surface meets all of these conditions ( Figure 6: Four-way valve for hydraulic
lift installation. Courtesy Weatherford International Ltd.). A high-pressure lubricator is placed on top
of the four-way valve for circulating the pump under pressure.

Figure 6

RECIPROCATING PUMPS: SYSTEM DESIGN


Hydraulic piston pumping involves seven basic design considerations:
1. Whether to use a closed power fluid (CPF) or an open power fluid (OPF) system.

2. Whether to pump produced gas with the liquid, or vent it up the casing or through a separate
tubing string.
3. What type of subsurface pump and downhole tubing arrangement to use.
4. What inflow performance characteristics should be expected from the well.
5. Whether to use a centralized power plant or wellsite power plant.
6. ipimsNumberedList >25. What size and type of power fluid surface pump to use.
7. What type of power fluid cleaning system to install.

POWER FLUID SYSTEM


A closed power fluid system tends to be more expensive than an open power fluid system
because it requires an extra flow conduit for the spent power fluid, along with higher-cost pumps
andwhere water is used as the power fluidthe addition of chemicals for lubrication, corrosion
inhibition and oxygen removal. It does, however, offer certain advantages:

The power fluid tank is relatively small. This makes the closed system a good option when
surface space is limited, as is the case at offshore or urban facilities.

Using water as a power fluid reduces the potential for fire or pollution in case of a leak.

If space and environmental considerations are significant, a closed power fluid system should be
selected. Otherwise, an open system is adequate.

PRODUCED GAS HANDLING


The decision of how to handle produced gas depends on the relative cost of the downhole
installation and the efficiency of the subsurface pump under producing conditions. The least
expensive downhole installations are those in which the gas and liquid production are comingled.
From this standpoint, it is better to pump the gas along with the liquid. But under certain conditions
for instance, if the well has a low bottomhole pressure and a high gas-liquid ratioproducing
the gas and liquids together can seriously reduce pump efficiency.
Consider, for example, a well producing 40 API oil with a 20% water cut and a gas-oil ratio of 500
SCF/Bbl, at a flowing bottomhole pressure of 400 psi. Pump efficiency under these conditions is a
very low 30 percent, as indicated in (Figure 1) (Hydraulic pump efficiency for various flowing
pressures, GORs and water cuts).

Figure 1

In general, if free gas production results in unacceptably low pump efficiencies (i.e., 30%-50%),
the gas should be vented. An alternative is to determine whether the desired production rate can
be achieved by increasing the bottomhole pressure, thereby holding more gas in solution. The IPR
must be reviewed in order to make this determination.

SUBSURFACE PUMP SELECTION


The next step in the system design is to consult the pump manufacturers catalogs to select a
subsurface pump that matches the wells expected inflow performance. Some pumps have single
engine-ends and single pump-ends; others have tandem engines with single pumps; still others
have tandem pumps with single engines, or tandem pumps with tandem engines.

PUMP SPECIFICATIONS
To illustrate the performance specifications for downhole hydraulic pumps, we consider the Kobe
Type A pump, which has a single engine and single piston ( Figure 2: Kobe Type A pump.
Courtesy Weatherford International Inc).

Figure 2

Table 1 lists specifications for pump sizes that fit in 2 7/8-inch O.D. or larger tubing (Weatherford
International Ltd., 2002-2004). Similar specifications are available for different tubing sizes and
pump models
Table 1
Specifications: Kobe Type A Downhole Pump
(Pump Sizes for Tubing O.D. = 2.875 inches or larger)
( after Weatherford International Ltd., 2002-2004)
Displacement
Pump Size

Area
Ratio
(P/E)

Rated
Spee
d,
SPM

Maximum Fluid
Lift, ft

Pump
End
at
Rated
Spee
d,
B/D

Engine,
B/D/SP
M

Pump,
B/D/SP
M

2-1/2 X 1 - -1

256

2.66

2.56

1.000

100

10,000

2-1/2 X 1-1/4 -1

256

5.02

2.56

0.520

100

15,000

2-1/2 X 1-1/4 -

367

5.02

3.67

0.746

100

13,400

-1-1/8
2-1/2 X 1-1/4 -1-1/4

492

5.02

4.92

1.000

100

10,000

2-1/2 X 1-1/4 -1-7/16

703

5.02

7.03

1.431

100

7000

2-1/2 X 1-7/16 -1-1/8

367

7.13

3.67

0.522

100

15,000

2-1/2 X 1-7/16 -1-1/4

492

7.13

4.92

0.700

100

14,300

2-1/2 X 1-7/16 -1-7/16

703

7.13

7.03

1.000

100

10,000

2-1/2 X 1-1/2 -1-1/2

745

7.55

7.45

1.000

100

10,000

2-1/2 X 1-5/8 -1-1/4

492

9.27

4.92

0.521

100

15,000

2-1/2 X 1-5/8 -1-7/16

703

9.27

7.03

0.770

100

13,000

2-1/2 X 1-5/8 -1-1/2

745

9.27

7.45

0.820

100

12,200

2-1/2 X 1-5/8 -1-5/8

944

9.27

9.09

1.000

100

10,000

2-1/2 X 1-7/16 -1-1/4 X 1-1/4

984

7.13

9.84

1.400

100

7200

2-1/2 X 1-7/16 -1-7/16 X 1-1/4

1195

7.13

11.95

1.701

100

5900

2-1/2 X 1-7/16 -1-7/16 X 1-7/16

1406

7.13

14.06

2.000

100

5000

2-1/2 X 1-5/8 -1-5/8 X 1-5/8

1818

9.27

18.18

2.000

100

5000

We can read these specifications as follows:

Pump Size: The first number refers to the nominal diameter of the pumpin this case it is 2.5
inches. The second number (in this case, ranging from 1 to 1 5/8 inches) indicates the
engine piston diameter. while the third number shows the pump piston diameter(s).
The first pump listed in Table 1 has a nominal diameter of 2.5 inches, an engine piston
diameter of 1 inch, and a pump piston diameter of 1 inch. Remember that the nominal
diameter indirectly specifies the tubing diameter. In this case, the nominal pump diameter of
2.5 inches will fit in 2-7/8-inch OD tubing.

Displacement: The displacement is expressed in terms of pump displacement at 100%


efficiency and maximum rated speed, engine displacement in B/D/SPM and pump
displacement in B/D/SPM. Thus, the sixth pump listed in Table 1 (2-1/2 X 1-7/16 - -1-1/8 )
has a total displacement of 367 B/D, an engine-end displacement of 7.13 B/D/ SPM and a

pump-end displacement of 3.67 B/D/ SPM. If both ends of this pump are operating at 100%
displacement efficiency, approximately twice the volume of power fluid is needed to pump a
given volume of produced fluid.
P/E ratio: This ratio varies according to the pump size, type and design features. It is equal to
the net cross-sectional area of the piston in the pump-end divided by the net cross-sectional
area of the piston in the engine-end. The P/E ratio is used as a design factor to determine
the required volume and pressure of the power fluid. The maximum P/E is determined by the
design of the power fluid system.Generally, the surface pressure of a power fluid system is
limited to a maximum of 5000 psi. Under these conditions, we may use the following rule of
thumb equation (Brown,1980):

Maximum P/E ratio = 10,000 / Net lift, ft


For example, if the net lift of produced fluids is 7,000 ft, then:
Maximum P/E ratio = 10,000 / 7,000 = 1.43

(1)

Thus, if we were to select a pump with a P/E ratio of greater than


1.43, we would exceed the 5000-psi limitation on the power fluid surface operating pressure;
this would not be the right pump for this specific application. Bear in mind that the value of
P/E for a pump depends on the design of the pump. The maximum P/E for the example
installation is determined by the design of the power fluid system.
Rated speed: This is given in strokes per minute. Multiplying this value by the pump-end
displacement in B/D/ SPM, gives the displacement of the pump.
Maximum lift: This is given in feet of fluid.

ACTUAL PUMP DISPLACEMENT


The displacements given in Table 1 are based on 100% pump and engine efficiency. The actual
pump displacement rate will be less than this ideal value. We need to determine whether this
displacement is sufficient to pump fluids at the desired production rate based on the wells inflow
performance. To make this determination, we must calculate the actual pump-end and engine-end
displacement volumes.
We begin by estimating the pump-end and the engine-end efficiencies. When a pump is new, the
pump-end efficiency is typically 90% and the engine-end efficiency is about 95%. This is usually
the case, assuming that we are not pumping at a high GLR. As the pump becomes worn from use,
the pump-end efficiency (E p) can fall to 70% or less, and the engine-end efficiency (E e) can fall to
about 80%.
The actual pump rate (q), can be estimated from Equation 2:
q = E p (PD)

(2)

where E p= pump-end efficiency (fraction)


PD = pump-end displacement at 100% efficiency
If gas is being pumped, then the pump-end efficiency value determined from ( Figure 1) should be
incorporated into an overall pump-end efficiency.

PUMP SELECTION EXAMPLE


Consider a well that we wish to produce under the following conditions:

Target production rate, based on IPR evaluation: 450 B/D


Net lift: 7000 ft
Production tubing string: 2 7/8-inch OD
Power fluid surface operating pressure limit: 5000 psi

Which one of the 17 Kobe Type A pumps from Table 1 would you recommend for this well?
Assume an engine-end efficiency of 90 percent, and a pump-end efficiency of 80 percent.

Solution:
From Equation 2, the required pump displacement rate is
PD = q / E p = 450/0.8 = 563 B/D

Of the 17 pumps listed in Table 1, we can eliminate seven of them from consideration based
on their rated pump displacements, which are less than 563 B/D.

Using Equation 1, we estimate a maximum P/E ratio. This enables us to stay within the power
fluid surface operating pressure limit. Of the 10 pumps remaining, we can eliminate four
from consideration because they do not meet the 7000-ft lift requirement and/or they equal
or exceed the maximum P/E ratio:
2-1/2 X 1-1/4 - -1-7/16
P/E = 1.431
2-1/2 X 1-7/16 - -1-7/16 X 1-1/4

P/E = 1.701

2-1/2 X 1-7/16 - -1-7/16 X 1-7/16

P/E = 2.000

2-1/2 X 1-5/8 - -1-5/8 X 1-5/8

P/E = 2.000

This leaves six pumps for consideration (Table 2):


Table 2: Pump Selection ExampleInitial Evaluation
Size:

Displacement:

P/E
ratio:

Maximum
lift:

2-1/2 X 1-7/16 - -17/16

703 B/D

1.000

10,000 ft

2-1/2 X 1-1/2 - -1-1/2

745 B/D

1.000

10,000 ft

2-1/2 X 1-5/8 - -1-7/16

703 B/D

0.770

13,000 ft

2-1/2 X 1-5/8 - -1-1/2

745 B/D

0.820

12,200 ft

2-1/2 X 1-5/8 - -1-1/2

944 B/D

1.000

10,000 ft

2-1/2 X 1-7/16 - -1-1/4 X


1-1/4

984 B/D

1.400

7200 ft

In these circumstances, we normally would consider the pump with the lowest P/E ratio, because
that pump should require the lowest power fluid surface operating pressure. The best pump
selection, therefore, would be the third pump listed above:
2-1/2 X 1-5/8 - -1-7/16
This pump has a 1 5/8-inch engine-end piston diameter and a 1 7/16-inch pump-end piston
diameter. Its pump end displacement is 703 B/D, with a P/E ratio of 0.770 and a maximum lift
capacity of 13,000 ft.
The actual rate of production for this pump will be 562 B/D, which satisfies the wells operating
requirements and enables the pump to be operated at less than maximum speed.
The actual required pump speed will be

(3)
where S
= required pump speed, SPM
S max = maximum pump speed, SPM
In this example,
S = 100 x (450 / 562) = 80 SPM

POWER FLUID VOLUME


The power fluid system must provide power fluid at the engine-end of the pump at a rate sufficient
to attain the desired production rate. To estimate how much power fluid is needed, we may use the
relationship

(4)

where q pf = actual throughput of power fluid, B/D


ED = engine-end displacement of 100% efficiency, B/D per SPM
S = required pump speed, SPM
E e = engine-end efficiency, fraction
Continuing with the Pump Selection Example above:

We have established a production rate of 450 B/D, and chosen a pump with a 1 5/8-inch
engine-end piston diameter and a 1 7/16-inch pump-end piston diameter to operate at 80
SPM.
From Table 1, the engine-end displacement for this pump is 9.27 B/D/SPM at 100% efficiency.
The engine-end efficiency, as already noted, is 90%. Thus, the required volume of power
fluid is

This means that if we use this pump, we will need 824 B/D of power fluid to displace 450 B/D
of production.

SURFACE OPERATING PRESSURE


To determine the surface operating pressure (ps), we must perform a force balance calculation for
the hydraulic system.

OPEN POWER FLUID SYSTEM


(Figure 3) is a schematic of an open power fluid system showing the various terms that are
necessary to undertake a pressure analysis:

Figure 3

ps = surface operating pressure, psi


F1 = frictional pressure losses in power fluid tubing, psi
G1 = hydrostatic gradient in power tubing, psi/ft
p1 = inlet pressure at engine-end, psi
p2 = pressure at outlet of engine-end, psi
Fp = frictional pressure losses in pump, psi
p4 = inlet pressure at pump-end, psi
p3 = pressure at outlet of pump-end, psi
G3 = hydrostatic gradient in production tubing, psi/ft
F3 = frictional pressure loss in production tubing, psi
pFL = backpressure on produced fluids, psi
h1 = pump setting depth, ft
h4 = pump submergence (i.e., liquid level above pump), ft
G4 = hydrostatic gradient of liquid in well, psi/ft
From an inspection of this figure, we may observe the following relationships:
p1 = ps + h1G1 - F1

(5)

(6)

p2 = p 3

(7)

p2 = pFL + h1G3 + F3

(8)

p4 = h4G4

(Alternatively, p4 is the bottomhole pressure at the pump under operating conditions.)


To calculate the pressure at any point in this system, we must be able to relate p 1 to p2, and p3 to
p4. Again using the Kobe Type A pump as an illustration, we start by looking at the forces acting on
the pistons during the downstroke (Figure 4: Cross section of the engine- and the pump-ends of
the pump, with the area of each section and the pressure applied to each during both the upstroke
and downstroke).

Figure 4

The forces acting on the various parts during the downstroke may be calculated as follows:
p1AR + p1(AE - AR) - p2(AE - AR)

(9)

At the pump-end of the pump, because of the valve port system, the following downward forces
exist:
p4(AP - AR) - p3(AP - AR) - p1AR
2

where: AR = area of rod, in


AE = area of engine piston, in2
Ap = area of pump piston, in2

(10)

In this particular pump, the piston diameters are the same at both ends, and so the net areas are
the same. Also, because the rod has a hollow tube passing through it that connects to a balance
tube below the pump, the area of the rod at its lower extremity is exposed to pressure p 1; this
imposes an upward force equal to p1AR. To these pressure terms, we must also add the pump
friction pressure losses (Fp).
Combining Equations 9 and 10, we obtain

(11)

We observe that the ratio of areas is actually equal to the net pump-end area divided by the net
engine-end areain other words, the P/E ratio, which we substitute into the equation:

(12)

p1 - p2 - (p3 - p4) (P/E) - Fp = 0

For this Kobe Type A pump, the equation for the upstroke will be identical to Equation 12. The
magnitude of Fp will depend upon the type of pump being used, and the viscosities and densities
of the power fluid and produced fluid at pump depth.
If the power fluid is water, we may obtain its viscosity from (Figure 5) (Correlation for determining
power fluid viscositywater).

Figure 5

If the power fluid is crude oil, we may estimate its viscosity using ( Figure 6) or similar correlations.

Figure 6

Combining Equations 5, 6, 7, 8 and 12, we obtain the following expression for p s, the surface
pressure required for an OPF system:
ps = (h1G3 + F3 + pFL)(1 + P/E) - h4G4 (P/E) + Fp + F1 - h1G1

(13)

To illustrate how we determine the required surface pressure for the OPF, we may return to the
Pump Selection Example introduced earlier, where

q = 450 B/D
Pump: Kobe Type A, 2-1/2 X 1-5/8 - -1-7/16
Pump P/E = 0.770
Pump speed: 80 SPM
Power fluid rate = 824 B/D
Pump setting depth (h1) = 7000 ft

In addition to this information, we have been obtained the following data:

Oil gravity = 40 API (specific gravity = 0.825; gradient = 0.357 psi/ft)


Produced water gravity = 1.07 (gradient = 0.463 psi/ft)
Power fluid: produced oil (G1 = 0.357 psi/ft)
Fluid level above pump (h4) = 2000 ft
Backpressure (pFL) = 50 psi
Bottomhole temperature = 160 F
Tubing diameters: 2 7/8-inch power string, 2 3/8-inch production string

The gradient of the wellbore fluids (G4) is assumed to be the weighted average of the produced oil
and water, or [(0.9 x 0.357)+(0.1 x 0.463)] = 0.368 psi/ft.
Because we have an OPF system, we assume the pressure gradient of production, G 3, to be
equal to the weighted average of the power fluid gradient and the produced fluid gradient. The
power fluid, produced at the rate of 824 B/D, has a gradient of 0.357 psi/ft and the wells
production of 450 B/D has a gradient of 0.368 psi/ft. This yields a weighted average of 0.361 psi/ft.
Thus:
G1 = 0.357 psi/ft
G3 = 0.361 psi/ft
G4 = 0.368 psi/ft.
Substituting into Equation 13:
pS = 1495.6 + 1.77 F3 + F1 + Fp

(14)

We see that there are still three unknowns: F3, the pressure loss caused by friction in the
production tubing; F1, the pressure loss caused by friction in the power fluid tubing; and F p, the
frictional pressure losses in the subsurface pump.
From (Figure 6), we estimate the viscosity of the crude oil at downhole conditions to be 1.8
centistokes. For the Kobe Type A pump operating at 80 percent of its rated speed (Brown, 1980),
Fp 450 psi x (specific gravity of oil) = 450 x 0.825 = 371 psi
Correlations for calculating the frictional pressure losses during the vertical flow of power fluid in
the power fluid tubing and the production in either tubing or annular conduits may be used to
estimate values for both F1 and F3 provided the GOR is low. If the GOR is high, two-phase flow
correlations should be used. With these correlations and information on the size of tubing or
annulus used for both power fluid and production, the rate of flow through each and the viscosity
of the fluids flowing, we may calculate both frictional pressure loss terms.
In this case, we assume that we obtain friction pressure losses of 16.8 psi and 84 psi from
published flow correlations.
With these values, we can return to Eq. 14 to calculate the power fluid surface operating pressure,
pS .

pS = 1495.6 + 1.77(84) + 16.8 + 371 = 2032 psi


This is significantly less than the maximum recommended value of 5000 psi. We now know that for
our OPF system, the power fluid must be injected at a surface pressure of 2032 psi and at a rate
of 824 B/D.

CLOSED POWER FLUID SYSTEM


The surface operating pressure, ps, for the closed power fluid (CPF) system is calculated in a
manner similar to that for the OPF system.

A CPF system has separate lines for the power fluid injection, power fluid return and produced
fluids, as shown in the schematic of (Figure 7). The variables referenced in the Figure are defined
as follows:

Figure 7

ps = surface operating pressure, psi


F1 = frictional pressure losses in power fluid tubing, psi
G1 = fluid pressure gradient in power fluid tubing, psi/ft
pPR = surface pressure of power fluid return line, psi
F2 = frictional pressure losses in power fluid return line, psi
pFL = surface backpressure on produced fluids, psi
F3 = frictional pressure losses in production tubing, psi
G4 = fluid pressure gradient in production tubing, psi/ft
p1 = pressure at inlet to engine-end of pump, psi
p2 = pressure at outlet of pump-end, psi
p3 = pressure at outlet of engine-end, psi
p4 = pressure at inlet of pump-end, psi

h1 = pump setting depth, ft


h4 = pump submergence depth, ft
Fp = frictional pressure loss in pump, psi
The following equations describe the flowing system:

(15)

p1 = ps + h1G1 - F1
p2 = pPR + h1G1 + F2

(16)
(17)

p3 = pFL + h1G4 = F3
p4 = h4G4

(18)

Note that the pressure available to drive the engine is p 1, the total pressure that the engine must
discharge against is p2, the pump end is filled by p3, and discharges against p4.
To obtain an equation that relates these four different pressures, we must perform a force balance
calculation for the upstroke and downstroke. For a Kobe Type A pump, the force balance equation
developed for an OPF system, Equation 12, is applicable.
p1 - p2 - (p3 - p4) (P/E) - Fp = 0

(12)

Substituting Equations 15 through 18 into Equation 12 and solving for p S:


pS = F1 + F2 + pPR + Fp + [(h1 - h4) G4 + F3 + pFL] P/E

(19)

To illustrate the use of Equation 19 for a CPF system, we may use the same parameters that we
used above for in the Pump Selection Example for the OPF system:
Pump Data:
Power fluid rate = 824 B/D
Production rate = 450 B/D
Production = 10% water cut
Pump Speed = 80 SPM
P/E = 0.770
Fluid Data:
Oil (power fluid + production) = 40 API => fluid gradient of 0.357 psi/ft (G 1)
Water specific gravity = 1.07 => fluid gradient of 0.463 psi/ft
Produced fluids (10% water) = fluid gradient of 0.363 psi/ft (G 4)
Depths:
Depth to pump = 7,000 ft (h1)
Pump submergence = 2,000 ft (h4)
Pressures and Temperature:
Backpressure in power fluid = 25 psi (pPR)
Backpressure on production = 50 psi (pFL)
Bottomhole temperature = 160 F
Installation:
Power tubing: 2 7/8-inch OD
Production tubing: 1 1/2-inch OD
Substituting these data into Equation 19 yields:
pS = F1 + F2 + 25 + Fp + [(7000-2000)(0.368) + F3 + 50](0.77)

=F1 + F2 + Fp + 0.77F3 + 1455.3


Using the same correlations described for the OPF system, we can estimate the following friction
pressure losses:
F1 = 16.8 psi
F2 = 16.8 psi
Fp = 371 psi
Frictional pressure loss for 450 B/D (1.5-inch tubing):
Viscosity of oil (at reservoir temperature) = 1.8 centistokes
Viscosity of water (at reservoir temperature) = 0.4 centistokes
Average viscosity (production) = (1.8)(0.9) + (0.4)(0.l) = 1.66 centistokes
=> F3 = 6.8 psi/1000 ft = (6.8)(7000/1000) = 47.6 psi
Substituting these values gives:
pS = 12.6 + 12.6 + 371+ 0.770(47.6) + 1455.3 = 1888.2 psi
The 1888 psi surface pressure required for the CPF system is lower than the 2032 psi required at
the surface for the OPF system. However, the lower surface pressure is obtained at the expense
of a separate return line for the power fluid.

SURFACE PUMP REQUIREMENTS


To estimate the size of the surface pump needed to supply power fluid at the required pressure
and volume, we may use the following hydraulic horsepower equation:
hp = q pf p (1.7 x l0 -5)

(20)

where:
hp = pump hydraulic horsepower
q pf = power fluid throughput rate, B/D
p = pressure difference between the suction and discharge sides of the pump, psi
For the Pump Selection Example (Open Power Fluid system):
q pf = 824 B/D
p = 2032 - 14.7 = 2017.3
and
hp = (824)(2017.3)(1.7 x 10 -5) = 28.26 hp
Of course, the pump would be ordered in the available horsepower ratings to meet this need, plus
any additional horsepower required to overcome pressure losses in surface controls units. If the
surface pump horsepower is higher than the design conditions require, be sure that this does not
overload the subsurface pump. Such an overload may result from delivering power fluid to the
engine-end of the pump so that it operates at a speed greater than its rated maximum or at a
speed greater than the inflow performance of the well. Also, according to one rule of thumb for
viscous crudes, an overload may occur if the speed is more than about half of its rated speed.
Pumps are normally tested using a hydraulic fluid with a specific gravity of about 30 API. For
more viscous crudes, the rated pump speed, and therefore the capacity, must be reduced to
accommodate the reduced rate at which the fluid moves through the pump-end of the pump.
In some cases, rather than installing a surface pump for each well, a single pump may be
designed to handle the power fluid requirements for a group of wells. In such cases, the pump
design must satisfy the total power fluid volume requirements of all the wells and be able to deliver
that volume at a pressure equal to the highest pressure required by any single well, plus about
200 psi to account for the pressure drop which occurs through surface flow control valves and

other equipment. Here, for example, if 200 psi is added to the 2017.3 psi estimated pressure drop
across the pump, the revised power requirements would be 31.1 horsepower.

POWER FLUID SURFACE FACILITIES


The final component in the design of a hydraulic pumping system is the power fluid surface facility.
One major task is to determine the tank capacity needed to adequately clean the power fluid in the
OPF system before re-injection. Standard tanks are normally 24 feet tall, with capacities of 300,
750, 1500, and 3000 bbl. These sizes provide adequate settling times for power fluid throughput
volume requirements of 600, 1500, 3000, and 6000 B/D, respectively. If more than a single tank is
needed, two tanks may be connected in parallel.
For the Pump Design Example that we have used throughout this discussion, the required power
fluid volume is 824 B/D. In this case, a 750-bbl tank, with a throughput volume of 1500 B/D, could
be used.

RECIPROCATING PUMPS: OPERATING


CONSIDERATIONS
If properly designed and installed, a reciprocating hydraulic pump system should operate
effectively for many years. There are, however, a number of operating issues to consider.
In an open power fluid system, where the power fluid returns are commingled with production, it is
often difficult to accurately measure the volume of formation oil produced compared with the
volume of power fluid. A similar problem can occur in a closed power fluid system if the power fluid
is lost during use. In such a case, it may be difficult to make accurate measurements of fluid
losses and operational efficiency.
Another consideration is the control of the pump speed, which relates directly to the volume of
fluid produced. It is the injection rate of the power fluid, rather than its pressure, that controls the
pump speed. Thus, the power fluid should be injected at a rate sufficient to meet design
conditions. Injecting at a higher rate can cause overloading of the pump.
More than 75% of hydraulic pump failures occur in the engine-end of the pump, and most of these
are caused by dirty power fluid. For this reason, one of the prime methods of analyzing pump
system behavior is to observe the changes in the operating pressure of the power fluid system.
For example, observing pressure fluctuations will show whether the pump is stuck or stalled, or
has suffered some mechanical problem (i.e., not on seat). Pressure fluctuations will also show
whether there is a leak in the power tubing or production line, whether the fluid level is fluctuating,
and whether the quality of production is changing (i.e., more water, emulsion).
The lifting of viscous crudes with hydraulic pumping systems should be approached carefully.
Pumps are rated using a test fluid that has approximately 30 API gravity. As a rule-of-thumb, the
power oil viscosity should not exceed 200 centistokes, and the pumped crude should not exceed
1000 centistokes. Even under these conditions, viscous crudes should be produced with the pump
operating at about one-half its rated speed. This situation may require a larger pump to sustain a
given pump rate. The pump manufacturer can provide guidelines for specific operating conditions.

JET PUMPS: DESIGN AND OPERATION


The operating performance of a jet pump can be adjusted by combining nozzles and throats of
different areas. This section describes the relationship between nozzle and throat size, and its
affect on production.

NOZZLE AND THROAT COMBINATIONS


Larger-diameter nozzles and throats will flow larger volumes. However, it is not the size of the
nozzle or throat but the ratio of their areas that determines the relationship of production pressure
and flow rate . Table 1 shows the qualitative performance characteristics that would result from two
different nozzle/throat area ratios.

Table 1: Qualitative Performance Characteristics, Jet Pump


(After Petrie, 1987)
Nozzle
Area, % of
Throat Area

Resulting
flow

Area
available for
well fluid
entry

Resulting production
rate, pressure head

Applicability

60

High
pressure,
low
velocity

Small

Low rate compared with


power fluid rate; high
head

Deep wells,
high fluid lift

20

Low
pressure,
high
velocity

Large

High rate compared with


power fluid rate; low
head

Shallow
wells, low
fluid lift

Various nozzle and throat combinations are used to achieve different inflow and lift height
requirements. Using a pump with a nozzle/throat-area ratio of 20 percent to lift small volumes of
well fluid would be inefficient due to pressure losses from highly turbulent mixing of the high
velocity power fluid jet and the slower moving well fluid. Conversely, attempting to pump large
volumes of well fluid with a nozzle/throat-area ratio of 60 percent would be inefficient because of
friction pressure losses caused by the produced fluid moving rapidly through the small throat. The
optimal ratios are those that balance the losses caused by mixing and by friction.
To achieve the best combination of nozzles, throats, injection rates, and well inflow, the optimum
nozzle size and ratio must be determined. A jet pump could be designed by trial testing each
combination of nozzles and throats to yield the best flow rates; a more efficient way is to calculate
the best configuration of nozzles and throats. Jet pump performance curves are useful for this
analysis. For a given nozzle, different nozzle pressures and throat sizes yield different
performance curves. Jet pumps are also sensitive to variations in intake or discharge pressure,
mixed fluid density and viscosity, and gas/liquid ratios. The calculations to simulate performance
with all these variables are difficult to solve without a computer.

CAVITATION
Another design consideration is cavitation, which occurs when the pressure at the head of the
throat is lower than the vapor pressure of the fluid. Cavitation will cause severe damage to the
pump. Therefore, anticipating and controlling the cavitation point is important. Cavitation is
prevented by maintaining the proper ratio of flow through the nozzle to flow through the throat.
This critical minimum flow ratio value can be computed for a combination of nozzles and throats.
Operating pressures and cavitation limitations often limit the number of possible combinations of
nozzles and throats for given well conditions.

EFFECTS OF GOR
Solution gas and free gas can pose challenges to the operation of jet pumps. The presence of gas
in the fluid increases the potential for cavitation and affects the operating characteristics of the
pump and power fluid requirements. The composition of well fluids must therefore be considered
when designing the jet pump system. By venting free gas, the effects on pump efficiency can be
reduced.

SYSTEM MONITORING
Jet pumps should be monitored to optimize well and pump performance. A comparison between
produced fluid rates and injected power fluid rates, will show whether the well's inflow performance

has changed. Drawdown and buildup data can be measured using downhole pressure recorders.
Producing bottomhole pressures can also be measured at different rates. Pressure recorders can
be set below the jet pump and the standing valve, and because a jet pump does not pulse, the
pressure recordings are very smooth.

Electrical Submersible Pump (ESP) Systems


ESP SYSTEM OVERVIEW
The major components of an Electrical Submersible Pump (ESP) system are shown in ( Figure 1).

Figure 1

The systems surface equipment includes transformers, a switchboard, junction box and surface
power cables. Power passes through a cable running from the transformer to the switchboard and
junction box, then to the wellhead
The ESP downhole assembly is located in the well at the bottom of the tubing. The motor, seal,
intake and pump assembly, along with the power cable, goes in the well as the tubing is run. The
well power cable is spliced to a motor cable that is connected to the outside of the downhole
assembly. Power from the surface facilities connects to the ESP at the junction box. The ESP itself
is a centrifugal pump located at the top of the downhole assembly (Figure 2: ESP systempump
diagram. Courtesy of Schlumberger).

Figure 2

Below the pump is an intake that allows fluid to enter the pump. Below the intake is a gas
separator and a protector or seal, which equalizes internal and external pressures and protects
the motor from well fluids. At the bottom is a motor that drives the pump. The assembly is
positioned in the well above the perforations; this allows fluid entering the intake to flow past the
motor and cool it.

ESP BENEFITS AND LIMITATIONS


ESP systems offer several advantages over other types of artificial lift:

They can be economically designed for both oil and water wells, at production rates ranging
from 200 to 40,000 B/D and at depths of .up to 15,000 feet.

They can be used in crooked or deviated wells.

They have a relatively small surface footprint, and so are appropriate for use in offshore,

urban or other confined locations.


Once in place, they are relatively simple to operate.
They generally provide low lifting costs for high fluid volumes.
They make it easy to apply corrosion and scale treatments.
Where justified economically (e.g., offshore locations where well intervention would require a
platform drilling rig or mobile offshore drilling unit), dual ESPs can be run in a single well; the
second ESP serves as a backup in case the first one fails (Duffy et. al, 2005)

Disadvantages of ESP systems include the following:

They are generally limited to single-zone completions


They requires a source of high-voltage electric power
Changes in well productivity may require expensive changes in downhole equipment.
The presence of a power cable alongside the tubing string can make it more difficult to run or
pull tubing.
They are not particularly good at handling gas and solids production.
Analyzing the system performance can be a challenge.
Power cables may deteriorate in high temperature conditions400 degrees Fahrenheit (about
200 degrees Celsius) is their general upper limit with respect to operating temperature.

ESP POWER COMPONENTS


This section describes the major components that supply power from the surface down to the
submersible pump.

TRANSFORMER
The first component of the ESP system is the transformer system, which is used to step-up or
step-down the voltage from the primary line to the motor of the submersible pump. The system is
usually arranged in banks of three single-phase transformers, a three-phase standard transformer,
or a three-phase autotransformer. Because a range of operating voltages may be used for
submersible pump motors, the transformer must be compatible with the selection of the motor
voltage.
For example, a primary voltage of 12,500 volts (V) may need to be stepped down to 2,400V by a
transformer. On the other hand, the primary voltage delivered to a well may be 440V, so a
transformer would be needed to step up that voltage to 880V. In general, the manufacturers of
electrical submersible pumps also build and sell the necessary transformers. An example of
transformer specifications, adapted from one page of a manufacturers catalog, is given in Table 1.
Table 1: Transformer SpecificationsSingle-Phase, OISC 60 Hertz
55 Degrees Centigrade Rise
(after TRW)
Size,
kVA

Height,
in.

Width,
in.

Depth,
in.

Weight,
lbs.

Delta
Primary
Volts

Delta
Secondar
y Volts

25

40

22.5

24.8

460

12500

480 / 960

25

47.5

22.5

28.8

500

24950

480 / 960

50

51.5

29

33

915

12500

650 / 1300

50

52.5

29

33

935

24950

650 / 1300

75

53.5

29

33

1095

12500

650 / 1300

75

59.5

29

33

1100

24950

650 / 1300

100

56.5

29

33

1325

12500

650 / 1300

100

62.5

29

33

1350

24950

650 / 1300

SWITCHBOARD
The electrical cable from the transformer goes to the switchboard ( Figure 1: Switchboard for ESP
system. Courtesy Weatherford International Ltd).

Figure 1

Switchboards sold by the ESP manufacturer are available in a range of sizes and models to
accommodate the electrical submersible pump system (Table 2)
Table 2: SwitchboardsGeneral Data
(After TRW)
Clas
s

Typ
e

Siz
e

Max
.
Volt
s

HP

Max
.
Full
Loa
d
Am
ps

Heigh
t,
inche
s

Width
,
inche
s

Depth
,
inche
s

Weight
, lbs.

DFH2

72

600

25

50

36.5

22.0

8.3

130

600

50

100

46.5

23.0

8.3

180

600

100

150

56.5

25.0

8.9

262

600

200

270

71.5

30.0

13.0

600

100
0

70

45

37.0

34.0

12.0

175

100
0

160

120

68.0

26.3

20.3

530

150
0

150

100

68.0

26.3

20.3

530

150
0

250

150

68.0

26.3

20.3

530

RPR2M

275
0

700

140

58.0

34.0

35.0

1100

RPR2S

275
0

700

165

58.0

34.0

35.0

1100

390
0

125
0

165

58.0

34.0

35.0

1100

490
0

125
0

165

58.0

34.0

35.0

1100

240
0

700

140

68.6

39.0

38.6

1095

275
0

700

165

68.6

39.0

38.6

1095

390
0

125
0

165

68.6

39.0

38.6

1095

490
0

125
0

165

68.6

39.0

38.6

1095

MFH

MDF
H

1512

76

76

76

The switchboard controls the pump motor and provides overload and underload protection.
Protection against overload (a condition where excessive amperage flows through the motor) is
needed to keep the motor windings from burning. Protection during underload (a condition where
the pump is not displacing its design volumes) is needed because low fluid flow rates will prevent
adequate cooling of the motor.
In addition to these functions, the switchboard may be used to record amperage on a continuous
basis, using a 24-hour chart. This chart is a good diagnostic tool for measuring pump
performance. The switchboard can also be used as an adjustable time-automatic restart control. In
this case, a pump that shuts down because the well is pumped off would depend on a switchboard
control to begin pumping again after a fixed time period. This protection is needed because the
pump should not be restarted until the previously pumped fluid has stopped backflowing through
the pump. Such backflow causes the motor impeller to reverse circulate. Placing a standing valve
in the tubing will normally eliminate this problem.
The switchboard may include additional features such as signal lights and automatic remote
control. Switchboards are available in ranges from 240 to 4800V.

JUNCTION BOX
The next component of the ESP system is the junction box, which connects the power cable from
the switchboard to the power cable from the well. As such, it should provide an explosion-free vent
to the atmosphere for any gas that might migrate up the power cable from the wellbore. The
junction box should be located at least 15 feet from the wellhead, and should be securely locked
at all times to protect against vandalism.

WELLHEAD AND POWER CABLE


A power cable, made up of three insulated conductors, runs from the junction box to the wellhead
(Figure 2: ESP power cable components. Courtesy of Schlumberger.)

Figure 2

A special wellhead is used to pack off the power cable so that it can enter the wellbore without
leaks . Cable is available in round and flat styles ( Figure 3: Round ESP cable; Figure 4: Flat ESP
cable). The conductors are available in different sizes and are usually made of copper or
aluminum.

Figure 3

Figure 4

The proper selection of the cable and the conductors depends on:

The expected amperage that will flow through the cable to the motor
The calculated voltage drop in the line from the surface to the pump.
The space that exists between the tubing collar and the casing (even though the cable is
banded to the tubing at selected points, there must be enough space to install and pull the
pump without damaging the cable or hanging it in the well).
The equipment operating environment - such as the operating pressure and temperature at
pump depth.

CABLE AMPERAGE
The first consideration in selecting cables is amperage. The limits on amperage for cables
containing copper conductors are as follows:
Cable No.

Maximum Amperage

115

95

70

55

Note that the cable with the smaller number has the larger diameter. Thus, a Number 1 cable can
carry a maximum of 115 amps.

VOLTAGE DROP
The second selection consideration is the voltage drop that will occur between the wellhead and
the pump. Normally, the maximum voltage drop for an electrical cable is about 30V per 1000 feet.
(Figure 5) is a graph of amperage versus voltage drop per thousand feet of cable length for
various conductors.

Figure 5

For example, if a 60-amp current is flowing through a 1000-foot cable, then the voltage drop in a
No.1 copper conductor will be about 16 V. In a No.4 copper conductor, it will be about 31V; while a
No. 6 copper conductor produces a voltage drop of about 45V. Clearly, the larger conductor with
the lower voltage drop is more desirable, especially in deep wells. A counter-argument to this is
that the larger cable costs more, and might not fit within accepted tolerances between the tubing
collars and the casing. These factors must be anticipated and balanced when designing a system.

ESP OPERATING PRINCIPLES


In most cases, submersible pumps are installed on tubing, although some are suspended at the
end of the power cable so that they can be retrieved without having to pull the tubing. In this
section, we will limit our discussion will to downhole pumps installed at the bottom of the tubing.
Electrical submersible pumps are multi-staged centrifugal pumps Each stage consists of a rotating
impeller and a stationary diffuser (Figure 1: Single stage of a centrifugal pump, showing impeller
and diffuser).

Figure 1

As the shaft of the pump moves in response to the force of the motor, the impeller turns, causing a
rotating motion in the fluid. The diffuser changes the direction and velocity of flow and directs fluid
from the impeller of one stage to the impeller of the next stage. The type of stage determines the
volume of fluid to be produced. The number of stages contained in a pump determines the total
pressure, or head, generated. The horsepower required by the motor is determined by both the
volume displaced and the head generated. Pumps are manufactured in a range of capacities to
satisfy almost all well conditions.
A pumps impellers are designed to operate efficiently over a specific capacity range ( Figure 2:
Optimal ESP operating range). Operating the pump below its design capacity causes the impeller
to downthrust against the diffuser, resulting in wear on the bearings and washers. Conversely, if
the pump operates above its design capacity, the impeller upthrusts against the upper part of the
diffuser, causing similar wear. Ideally, the impeller should float freely, and will do so throughout its
recommended operating range. This recommended operating range will allow the pump to run at
highest efficiency.

Figure 2

Pump length and diameter are constrained by manufacturing and wellbore conditions. Assembly
and handling difficulties usually limit the length of a single pump section to about 20 to 25 feet.
However, it is possible to join pump sections together, adding successive stages to develop the
required head. Pump diameters are limited by the size of casing in which they are to be run. Table
3 lists the suggested pump diameters for various casing sizes; for example, a 3 3/8-inch OD pump
will fit within 4 1/2-inch OD casing. If the casing diameter is 8 5/8 inches or greater, then larger
pumps could be used.
Table 3: Recommended ESP diameters for various casing sizes
(after Brown, 1980)
O.D. of
casing

Casing
weight

Motor O.D.

Pump O.D.

4.50 inch

11.5 lb/ft

3.75 inch

3.375 inch

9.5 lb/ft
5 inch

All weights

3.75 inch

3.375 inch

5.5 inch

20 lb/ft

4.5 inch

4.0 inch

17 lb/ft
15.5 lb/ft
7 inch

28 lb/ft

4.5 inch

4.0 inch

26 lb/ft

5.5 inch

5.375 inch

4.5 inch

4.0 inch

5.5 inch

5.125 inch

7.375 inch

6.75 inch

24 lb/ft
20 lb/ft
8.625 inch or
greater

All weights

The number of stages that can be added to a pump are limited by three variables:

Horsepower rating of the shaft that turns the pump


Pressure rating of the pump housing
Capacity of the thrust bearing

The pump manufacturer takes each of these variables into account when a customer specifies the
capacity range for a pump.

ESP SYSTEM DESIGN


This section examines the considerations and design procedures involved in selecting a pump and
motor for an ESP system, and the steps needed to determine the surface voltage requirements.

PUMP PERFORMANCE CURVES


In selecting a pump for a particular application, we must look carefully at its performance or test
curves, which typically chart three different aspects of performance: head versus pump capacity,
motor horsepower versus capacity and pump efficiency versus capacity ( Figure 1: ESP
performance curve. Courtesy Schlumberger-Reda.)

Figure 1

These curves are published by pump manufacturers for each of their individual pump typesthe
ones shown in (Figure 1), for example, are for a 100 stage, Schlumberger-Reda D1350 pump .
They are obtained by running a pump in water at a constant speed, while varying its throughput by
throttling the discharge side of the pump. During the test, the pressure difference across the pump,
the brake horsepower, and the pump efficiency are measured at different pump throughput rates.
The resulting pressure increase is then converted to its equivalent head. With this data,
performance curves are drawn showing head, pump efficiency, and brake horsepower for a
specified number of pump stages, as a function of pump throughput rate.
Although these curves are generated using fresh water (with a specific gravity of 1.0), the same
values of head are usually used when selecting a pump for a fluid with a different specific gravity
provided the viscosity of the fluid is similar to the viscosity of water. Brake horsepower, on the
other hand, does require a specific gravity correction.
Pump performance curves are normally published for either a single pump stage or, as was done
in Figure 1, for 100 pump stages.

PUMP HORSEPOWER REQUIREMENTS


Using the pump performance curve shown in (Figure 1) above, we may determine the
horsepower requirements for a pump under a given set of operating conditionsfor example, a
flow rate of 1250 B/D and a required pump pressure of 1725 psi.
First, we note the following values corresponding to a capacity of 1250 B/D:

Head generated for 100 stages = 2280 ft, or 22.8 ft per stage
Brake horsepower required = 32.5 hp, or 0.325 hp per stage
Pump efficiency = 64 percent.

The number of stages required is:

(1)
The total head is equal to the required pump pressure divided by the pressure gradient of fresh
water (the fluid upon which the pump curve is based), or 0.433 psi/ft. Therefore,

To calculate the required brake horsepower (hp), we multiply the horsepower per stage by the
number of stages and by the specific gravity of the fluid being pumped (in this case, the fluid is
fresh water, and so the specific gravity is 1.0):
hp = (hp per stage) x (stages) x (specific gravity)

(2)

= (0.325) x (l75) x (1.0) = 56.9 hp


In summary, this procedure requires specifying the pump operating conditions (required capacity,
in B/D, and the required fluid head). The required capacity of the pump is based on an estimate of
the wells inflow performance. The performance curves are actual test curves. If we assume that
very little gas will be pumped, we do not need to apply a pump efficiency factor. We should select
the pump to provide the desired capacity based on continuous operation.
In evaluating the capabilities of various pumps, we may have a number of options from which to
choose. It is important to study the performance curves supplied by the manufacturer to select a
pump that, at the required capacity, will operate in its optimal capacity range at the highest pump
efficiency. Remember that pump capacity depends on the design of the impeller and not on the
number of stages. With the selected performance curve and pump capacity, look up or calculate
the head/stage and horsepower/stage. Once we know the required head, we can calculate the
total number of stages needed and the horsepower required for the motor.

TOTAL DYNAMIC HEAD


Along with specifying the required capacity of a pump, we need to determine how much total
dynamic head (TDH) it must provide. The TDH is the total head required when pumping at the
desired rate. It is the difference between the head at the pump discharge and the head at the
pump intake.
Assuming the produced fluid has no free gas, and that all pressures are converted to head (in
feet) using the specific gravity of the produced fluid, TDH is equal to:
the head caused by backpressure in the tubing,
minus the frictional losses generated by flow in the tubing,

plus the head contributed by the liquid column in the tubing,


minus the head caused by the operating liquid column in the annulus,
minus any head caused by a backpressure imposed on the annulus.
Generally, the annulus backpressure is negligible and so the liquid columns may be netted out to
give a single term for the net lift. As a rule of thumb, a pump should have at least 500 feet of fluid
above it when in the operating mode.
As an example, consider a well producing under the following conditions:

Tubing backpressure is 75 psi.


Pump depth is 8500 ft; operating fluid load depth is 700 ft above the pump.
Tubing size is 2 7/8-inch OD; casing size is 7-inch OD.
Desired pump rate is 1400 B/D.
Fluid being pumped is 25-degree API oil (specific gravity = 0.904; gradient = 0.392 psi/ft).
Assume that there is no packer in the hole

Because there is no backpressure on the annulus, the TDH, expressed in feet, can be estimated
using the following equation:
TDH = head due to tubing backpressure + frictional losses + net lift

(3)

Tubing backpressure head =


The frictional losses during flow in the tubing are obtained using viscosity correlations provided by
the pump manufacturer, and are found to be approximately 9 psi per 1000 feet of tubing. This
converts to

The net lift, with no packer in the hole, is simply the depth to the pump, minus the height of the
fluid level above the pump:
Net Lift = 8500 - 700 = 7800 ft
When all values are substituted, the total dynamic head is:
TDH = 191.3 + 195 + 7800 = 8186.3 ft
This is the value of TDH to use when selecting a pump.

MOTOR SELECTION
Once we determine a pumps required capacity and TDH, the next step in the design is to review
performance curves from different manufacturers. The goal is to (1) select the pump that best
meets the capacity needs, (2) calculate the number of stages required, and (3) specify the
required motor horsepower.
To illustrate, assume the same well conditions that were given in the Total Dynamic Head
calculation above, and use the resulting TDH value of 8186 feet. Assume also that after reviewing
the performance curves, the pump from (Figure 1) was selected . This pump has an OD of 4
inches and will easily fit within the 7-inch casing.
The performance curves for this pump show that at a throughput rate of 1400 B/D and a TDH of
8186 ft:

head = 2100 ft / l00 stages


hp = 34 hp/100 stages
The number of stages and motor horsepower required are as follows:
stages = 8186 ft divided by 21 ft/stage = approximately 390 stages
hp = (hp / stage) x (stages) x (spec. grav.) = (0.34) (390) (0.904) = approximately 120 hp
If the bottomhole temperature is higher than 180 F, the horsepower required should be increased
by as much as 20%. In the example well, high temperature is not a limiting factor, so we can go
with 120 hp in our design.
Next, we need to select a 120 hp motor that will fit into a 7-inch OD casing. In this case, we can go
to our manufacturers catalog and see a listing of 120 HP Series 456 motors with the
specifications shown in Table 1.
Table 1: 456 Series Motors (4.56-inch OD)60 Hz
HP

Volts

Amps

120

1000

77

120

1170

66

120

1350

57

120

2300

34

The Series 456 has an OD of 4.56 inches, which easily fits inside the 7-inch casing. Each of the
120 HP motors listed above has a different required voltage; note that the higher voltage motors
require lower current.
Now we must select one motor for the application. In general, the choice of motor voltage is a
function of line voltage losses, capital costs, and the electric power cost.
The required surface voltage is the voltage required by the motor, plus voltage losses between the
motor and the surface, including losses in the cable, other system components, and the
transformer. The cable losses are the most significant. As an upper limit, cable losses should be
less than 30V/1000 ft.
If, for example, we select the 1000V, 77-amp motor from Table 1, we can then go to ( Figure 2)
(Amperage per voltage drop per 1000 feet of cable, various conductors).

Figure 2

We note that an ideal conductor is the No. 2 copper conductor, which at this amperage has a
voltage drop of 25V per l000 feet of cable.
Another option would be the 2300V, 34-amp motor. This would allow us to use a smaller No. 6
copper conductor and have the same voltage drop along the cable. Because the No. 6 copper
conductor is smaller than the No. 2, going with this higher-voltage option would result in a lower
capital cost for the cable. However, a higher voltage switchboard will be more expensive than a
lower voltage switchboard.
We could also consider using a larger conductor cable (for instance the No. 2 copper cable) with
the 2300V motor. The voltage loss in the cable will drop from 25V to 11V per 1000 ft. The larger
conductor cable will require a higher capital cost, but lower operating costs. Within limits, then, a
larger, more expensive cable will allow for lower line losses and thus lower operating costs.
In summary, motor selection requires an economic analysis of voltage and cable alternatives;
however, the following recommendations should serve as a guide:

For low horsepower motors and shallow depths, use a 440V system.
For less than 70 hp at intermediate depths, use an 830V system.
For 70 to 200 hp in deep wells, use a 1500V system.
For motors of 200 volts and greater, use a1500V or 2400V system. The choice of voltages will
depend on economics.

SURFACE VOLTAGE REQUIREMENTS

Before we can select a switchboard and transformer for our ESP system, we must determine the
required surface voltage, Vs:

(4)

V s = V m + V c + V com + V t
Where:
V s = required surface voltage, volts
V m = motor voltage, volts
V c = cable voltage losses, volts
V com = system component losses, volts
V t = transformer voltage losses, volts

Consider the 2300V motor for analysis. With the this motor, the No. 1 copper conductor, and a
pump depth of 8500 ft, the voltage loss in the cable, V c, is 76.5 volts. Assume the surface
component losses, V com, are negligible. The transformer losses, V t, are usually estimated to be
2.5 percent of the required voltage. In this case the required surface voltagewithout significant
electrical component voltage losses and using the No. 1 copper conductorwould be:
V s = 2300 + 76.5 + 0 + 0.025(2376.5) = 2436 V
Operating voltage is variable within a 50V range; thus, a standard 2400V transformer system
would meet this systems voltage requirements. In most cases, it is better to have a required
surface voltage below 2400V. For this reason, we should analyze the other motor and cable
options to find one that will provide a surface operating voltage somewhat below that of the
transformer system. Further analysis of this specific pump system shows that the 1170V motor
with a No. 4 copper conductor cable will require a surface voltage of 1495V. With this pump and
cable, we may specify a system voltage of 1500V.
The switchboard would have a maximum voltage rating of 1500V and a maximum amp load of at
least 66 amps.
The transformer size is expressed in kilovolt-amps, kVA, and is given by the expression:

(5)

kVA = 0.00173 V s A m
Where:
kVA = transformer size, kV-amps
V s = system voltage, V
A m = motor amperage, amp

For this example, V s is equal to 1495 volts and A m is equal to 66 amps. So the transformer size
is:
kVA = 0.00l73(l495)(66) = 171 KVA
If we use three single-phase transformers, each will require one-third of this value, or about 57
kVA. Other specifications of the transformer would be determined in consultation with the
manufacturer and would include consideration of the primary line voltage available in the operating
area (Table 2 ).

Table 2: Transformer SpecificationsSingle Phase OISC 60 Hz, 55


(After TRW)
Size, Height, Width, Depth, Weight, Delta Primary
kVA inches inches inches
lbs.
Volts

deg C Rise

Delta Secondary
Volts

25

40

22.5

24.8

460

12500

480 / 960

25

47.5

22.5

28.8

500

24950

480 / 960

50

51.5

29

33

915

12500

650 / 1300

50

52.5

29

33

935

24950

650 / 1300

75

53.5

29

33

1085

12500

650 / 1300

75

59.5

29

33

1100

24950

650 / 1300

100

56.5

29

33

1325

12500

650 / 1300

100

62.5

29

33

1350

24950

650 / 1300

PRESENCE OF EXCESSIVE FREE GAS


The TDH calculation outlined in this section is accurate if significant volumes of gas are not being
pumped, and if no gas is coming out of solution as fluid is produced up the tubing. If gas is
present, then additional considerations come into play.

Normally the pump is installed on the tubing without a packer. This means that free gas is

easily vented up the annulus. (Many ESP well installations--especially offshore--do include
packers.)
It is not feasible to vent annulus gas to the atmosphere so it must travel through surface flow
lines. However, the annulus backpressure and any surface venting must be controlled so
that the pump is below an acceptable level or fluid submergence. If not, the pump may
become gas-locked.
Manufacturers provide downhole gas separators to assist in preventing gas from entering the
pump.
If the pump is set at a depth at which the pressure is below the bubble point, any free gas that
is not separated will be pumped. The pump will perform normally if the volume of free gas is
below 10%.
If significant gas volumes are present in the fluid, a higher capacity pump will be needed. An
offsetting factor is that the presence of gas in the tubing will reduce the TDH and thus
reduce the required motor horsepower.

All other considerations equal, it is more difficult to size pumps for wells that pump gas, because it
becomes necessary to estimate the pressure-volume effects that occur between the pump intake
and discharge pressures. This is illustrated in ( Figure 3) (Effect of gas compression on ESP
throughput) .
Note that the initial pump stages perform a substantial amount of gas compression. Between an
intake pressure of 500 psi and a discharge pressure of 1215 psi, the fluid volume is reduced
almost by half.

Figure 3

Handling gas effects in the ESP design generally involves the following steps:
1. Estimate the pump intake pressure from inflow data
2. Estimate the pump discharge pressure from multiphase flow correlations
3. Divide the range of pressure between the inlet and discharge valves into discrete increments
and then calculate the pressure-volume changes that occur between increments
4. Calculate the average volume throughput and average fluid gradient, from the pump intake to
discharge points, and, with this data, select a pump and then specify the number of stages
required using the pumps performance curves.

The appropriate calculations are typically incorporated into company and manufacturer ESP
design software. You should check on the individual program that you are using to ensure that gas
effects are taken into account where appropriate.

ESP SYSTEM OPERATION


One disadvantage of ESP systems is that the tubing string normally has to be pulled in order to
replace the pump or other subsurface equipment (unlike, for example, a free hydraulic pump, a
rod pump or a wireline-retrievable gas lift valve). This requires a workover rig, which adds to the
cost of well maintenance and repair.
Replacement of downhole equipment may become necessary because of failures at the motor,
pump or cable (Table 1), or because the pump needs to be re-sized due to changes in the wells
inflow performance.
Table 1: Common ESP Failures
Type of
failure

Possible causes

Motor

Excessive motor overload


Leak in the protector caused by worn or defective seals
Insufficient fluid movement to cool the motor
Motor housing corrosion
Faulty installation
Switchboard problems
Lightning

Unbalanced electrical system


Pump
section

Downthrust or upthrust wear on the pump bearings


Wear caused internally by the pumping abrasives
Plugged or obstructed stages caused by deposition or corrosion
Twisted shaft caused by starting a pump while the fluids are still moving
downward in the pump after shut-down

Cable

Mechanical damage caused during running or pulling operations


Deterioration caused by the downhole environment
High temperatures caused by amperage loads

If the wells inflow has been increasing, the cost of bringing in a workover rig is relatively easy to
justify, because the newly installed pump will pay for itself with the higher production rate. But if
the inflow is decreasing, the expense is harder to bear, and the temptation might be to just cycle
the pump daily instead of replacing it.
The problem with cycling an ESP is that large current surges occur when starting up a high
voltage, high horsepower induction motor that is separated from its starter at the surface by five or
ten thousand feet of cable. These surges are often five to eight times the normal operating current
(in Figure 1, a normal 32-amp current surged to 141 amps at start-upNeely, 1982). This type of
starting is referred to as across-the-line starting. Starting and stopping the motor also causes
amperage and voltage spikes that can damage or burn out the motor. Normally, only a limited
number of stops and starts are possible, so most operators are reluctant to stop and start a
submersible pump unless it is absolutely necessary.

Figure 1

OPTIMIZING PUMP PERFORMANCE


There are several ways to optimize pump performance as the well inflow performance changes.

CHANGES TO THE MOTOR AND PUMP


If well inflow performance is decreasing and the pump is oversized and continuously running,
pump performance can be improved by:

Changing the pump. This requires pulling the tubing, losing some production, and buying and
installing a new pump. Try to avoid this option.

Increasing the backpressure at the wellhead. This will reduce the capacity of the pump. This is
also undesirable because of the energy waste and the potential for thrust problems in the
pump.

The remaining alternatives require modifying the electrical submersible pump system.

Cycle the pump and motor. This method eliminates the current surges during start and stop
operations. This is similar to what is done with the motor of a rod pumping system.

Reduce the motor speed and, thus, the speed of the pump. This is the preferred method of
changing the pump capacity.

SOFT START/SOFT STOP


The surges in current during start and stop operations may be limited by using silicone-controlled
rectifiers at the surface starter to provide controlled levels of power at the pump. With these
rectifiers, the voltage and amperage are allowed to ramp up or ramp down at acceptable rates
over a specified period of time. This type of start and stop control, which provides smooth power
flow to and from the motor during stop/start operations, is referred to as soft start and soft stop
(Figure 2). Soft-start controllers have been used under field conditions and have demonstrated
that pumps may be cycled numerous times without damage (Neely, 1982).

Figure 2

VARIABLE FREQUENCY DRIVE


Another option for changing the capacity of an electric submersible pump is to install a surface
device that can change the voltage frequency supplied to the motor. This device is referred to as
a variable frequency generator (VFG), variable frequency drive (VFD) or variable-speed drive
(VSD). It is used instead of the standard 50- or 60-Hz motor controller at the surface.
As an example, a VFG may be rated at 300 KVA with a frequency range of 36 to 90 Hz. Because
the motor speed is proportional to frequency, and the pump speed is equal to the motor speed, a
variable speed generator allows the pump to operate at a speed above or below its rated capacity
at 60 or 50 Hz. There is an upper limit, though, on pump speed variability. This limit is at the
amperage overload condition of the downhole motor. Subject to this limit, a variable frequency
generator can be used to increase or decrease pump rates by as much as 20% to 30%, to satisfy
changing inflow performance. If greater changes are required, a resizing of the pump will be
necessary.

Variable frequency generators and variable-speed drives can also provide soft-start protection.
Using a VSD to control a pump motor can reduce the strain on the pump shaft and reduce pump
damage by slowly ramping-up the motor speed. A number of operating procedures have been
developed to make the electrical submersible pump more versatile in its application to a specific
well or field. However, the best approach is to carefully design and select a pump for each well
based on its own characteristics and assume that the pump will operate continuously.

ESP SYSTEM MONITORING AND CONTROL


ESP systems are complex and require specific operating conditions with respect to the power, fluid
level, fluid content, and gas-liquid ratios. Without proper ESP monitoring, operators cannot tell
what is happening downhole, and this lack of monitoring may lead to low efficiency, high lifting
costs and frequent repair and replacement. Poor ESP performance can also result from improperly
designed pump assemblies. Monitoring and performance analysis can extend the life of ESP
pump installations by reducing system down time and the cost of repair and replacement of the
pumps. Reducing down time also results in higher production from the well, improved ESP
performance and increased profitability.
Analysis of ESP performance collected from the wellsite data gives the operator real-time
information, historical data, and the ability to optimize the wells from a remote location. Monitoring
and comparing changes in ESP performance over time can help identify potential problems and
correct them before the ESP fails.
Typical components include:

Downhole pressure, temperature, and electrical sensors


Surface System using well sensor data
Remote Terminal Unit (RTU) and communications system
Interpretation and Analysis Software

DOWNHOLE PRESSURE, TEMPERATURE, AND ELECTRICAL


SENSORS
A well controller gathers data from downhole sensors that provide measurements of ESP
performance. The sensors can also measure data such as pump intake pressure, wellhead
pressure, and flow rate. Data of this kind can help identify problems before they become more
serious. Pump wear, paraffin or scale buildup, sand production, pump off, and blocked perforations
can also be identified by making downhole measurements of changes in current and voltage,
power factors, and pump intake pressure.

SURFACE SYSTEMS USING WELL SENSOR DATA


An ESP surface controller may be connected to the downhole sensors and surface instruments to
measure conditions and control the operation of the ESP pump in each well. The operator may
want to monitor surface performance parameters such as wellhead pressure, net oil, flow volume,
and flowline temperature, among others.
The controller can be programmed to automatically adjust pump operations using data from the
well sensors. The controller regulates the cycling operations of well so that startup of the well is
smooth, reducing stress on the motor and pump, improving pump efficiency, and extending the run
life of the system. The controllers can also be programmed to send data from the downhole
sensors, motor controller to the remote locations using the RTU.

REMOTE TERMINAL UNIT (RTU) AND COMMUNICATIONS


SYSTEM

The sensor data must be communicated from the well to the operator. A typical ESP monitoring
system has a remote terminal unit that sends the sensor data to the operator. The RTU may use
radio or telephone modem to transmit wellsite data to the operators office.

INTERPRETATION AND ANALYSIS SOFTWARE


The software, running on a desktop platform, allows the operator to monitor and set the ESP
controller on each well in a field.
The ESP monitoring and analysis software allows the operator to control, monitor, track, and
report on ESP wells. It provides a means of reporting production parameters, tracking lifting costs
and detecting operating problems. Proper monitoring of ESP performance can increase pump life,
reduce operating and repair costs and improve profitability.
With the analysis of pump performance data, the operator can model reservoir inflow and pump
performance so that the pump can run at its highest efficiency. The operator can also simulate the
performance of several combinations of pump size, setting depth and production rates. This allows
the operator to avoid designs that cause stress on the pump, motor, or cable.

ESP SYSTEM DESIGN USING FIELD DATA


Observations of ESP performance over a long time are used to design the best pump
configuration by modeling various combinations of pumping system equipment. The operator can
experiment with parameters such as pump setting depth, production rate, and fluid level to
determine how these changes affect well performance.

Plunger Lift Systems


PLUNGER LIFT OVERVIEW
Plunger lift (Figure 1) is the only artificial lift method that relies solely on the wells natural energy
to lift fluids. The plunger, traveling inside the tubing, moves upward when the pressure of the gas
below it is greater than the pressure of the liquid above it. As the plunger travels to the surface, it
creates a solid interface between the lifted gas below and produced fluid above to maximize lifting
energy. Any gas that bypasses the plunger during the lifting cycle flows up the production tubing
and sweeps the area to minimize liquid fallback.

Figure 1

The plunger is a cylinder of slightly smaller diameter than the tubing diameter. The exterior of the
plunger can have different types of packing elements to seal against the tubing wall. Plungers may
be solid or they may have an internal rod-actuated bypass valve.
In addition to the plunger itself, a typical assembly consists of two bumper springsone in the
bottom section of the tubing, and the other in a lubricator at the surfaceand a surface controller
to control the travel cycle of the plunger.
The plunger cycle consists of three stages (Phillips, 1998):
1. Shut-in: During this stage, the well is shut-in to allow the casing pressure to build to the level
required to lift the plunger and the liquid above it.
2. Unloading: The tubing is opened, and the stored casing pressure lifts the liquid and the
plunger to surface.
3. Afterflow: With the plunger at the surface, the well is allowed to flow in preparation for the next
shut-in period At the end of the afterflow period, the well is shut-in, the plunger falls to
bottom, and the casing pressure again begins to build.

Plunger lift provides a cost-effective method of artificial lift that can be used to efficiently produce
both gas wells with fluid loads and high GOR oil wells. It is the only artificial lift method that uses
the natural energy of the well to lift fluids. It can be used to

APPLICATIONS

Plunger lift has found successful application in a variety of production scenarios, including:

Producing high-GOR oil wells


Increasing production from oil wells with low bottomhole flowing pressures
Dewatering gas wells or wells in coalbed methane reservoirs
Unloading of wells being converted to gas lift
Reducing liquid fallback in intermittent gas lift wells
Cleaning wells with paraffin, scale or
Improving production in wells with emulsion problems
Maintaining production in wells that are awaiting other permanent artificial lift process
Producing single-well properties or small fields

Advantages

Can be used in wells as deep as 10,000 ft TVD.


Given a sufficient producing GLR, plunger lift operations require no external gas supply.
Because there are few moving parts, the system is easy to service and maintain.
Reduces need for hot-oiling or mechanical paraffin cutting
Works well in deviated wells (if the plunger operates in a relatively straight part of the well)

Disadvantages
At low velocities (350-400 ft/min), the plunger may stall in the tubing.
High velocities (1200-1500 ft/min) may reduce fluid recovery (more fluid fallback).
The plunger also may be damaged if it hits the upper bumper spring too fast.

(Optimum plunger velocity is 800 to 1,000 ft/min. The plunger velocity can be estimated by
measuring how long it takes the plunger to arrive at the surface after the time-cycle controller
opens and gas injection begins. Average plunger velocity can be determined by calculating the
depth of the well, divided by the travel time.)

TYPES OF PLUNGER INSTALLATIONS


Conventional plunger lift without a packer is typically used in wells that have enough formation gas
to lift the fluid. This is the most common type of plunger lift installation.
Intermittent gas lift using a plunger is for wells with low bottomhole pressure. In this case, the
height of the fluid column allows gas to break though during the lift cycle. The plunger keeps the
gas and fluid separated, thus reducing fall back while increasing fluid inflow. The gas required for
moving the plunger is injected from the surface.
Plunger lift with a packer is for gas wells or high-GLR oil wells, and uses a packer set in the
tubing-casing annulus. When the plunger reaches the surface, it is held there to allow fluid flow to
the sales line. In these wells, timers normally control the injection cycle so that the plunger will
remain at the surface for a sufficient length of time.

PLUNGER LIFT DESIGN AND INSTALLATION


Plunger lift can be easily installed in a well using wireline, so it is usually not necessary to pull the
tubing. A typical plunger installation is shown in (Figure 1) (Courtesy Weatherford International
Ltd).

Figure 1

To stop the plunger when it falls to the bottom, a tubing stop and bumper spring assembly are set
by wireline at the bottom of the tubing string. This bumper spring is positioned just above the
perforations. In wells with low permeability producing formations, a standing valve may be installed
below the perforations, between the tubing stop and the bumper spring.
At the surface, a lubricator above the master valve contains:

a flow outlet,
a bumper spring and striker pad to stop the moving plunger,
and a catcher to hold the plunger so it can be removed for service.

A device to detect the arrival of the plunger is also part of the control system. The arrival of the
plunger at the surface closes the gas injection control valve.

MECHANICAL EQUIPMENT
The main components of a plunger lift system are the plunger, bumper springs, lubricator,
controller and the surface gas injection facility.
Plungers are manufactured in various diameters, lengths, and temperature ratings to
accommodate specific fluid compositions and well conditions. Plungers may be solid or may have
a rod-actuated internal bypass valve. Solid plungers fall back to the bottom slowly and are used in
wells with longer cycle times. Plungers with a bypass valve are used in wells with shorter cycle
times where it is important for the plunger to fall to the bottom more quickly. When the plunger is
on bottom, the bypass valve is closed, to keep gas from flowing through it. When the plunger
arrives at the surface, the valve rod strikes a bumper spring and pad in the lubricator, thus
opening the valve and allowing it to fall to the bottom. At the bottom, the rod contacts the lower
bumper spring and pad and closes the valve for the next injection cycle.

Brush plungers have a nylon brush on the outside to move fluid, fine particles, and sand through
the tubing. Spiral plungers have hard concentric rings around the outside of the plunger body and
work better than brushes in high GLR wells and in wells with paraffin or scale. Expanding blade, or
pad, plungers use expansion sections that are held against the tubing wall with springs. The
spring-mounted pads follow any tubing irregularities and maintain the seal. Other plungers use
combinations of brushes, spirals and pads.
In normal operations, the plunger travels in the tubing between a bumper spring at the bottom of
the tubing and the lubricator and a bumper spring above the wellhead.
At the start of a lift cycle, the plunger falls to the bottom of the tubing. Well fluids accumulate in the
tubing above the plunger. The surface controller causes gas to be injected into the wellbore below
the plunger through perforations in the casing or through gas lift valves. The plunger lifts the fluid
to the surface when the force of the injected gas beneath the plunger exceeds the weight of the
fluid above it. The plunger keeps the fluids separated from the gas, reducing fluid fallback. The
plunger travels into the lubricator where it is held until it is time for the plunger to fall again to the
bottom. The lubricator contains a sensor to detect the arrival of the plunger at the surface.
Surface controllers regulate the frequency of the plunger movement, and these are adjusted to
match the well's inflow performance. The surface controllers are activated by a pressure device or
a timer. Pressure controllers are often used in oil wells with low GLRs or in wells with high PI's,
whereas controllers using timers are frequently used on gas wells and high GLR oil wells. An
example of a modern controller system is one that provides time- or pressure-based (or a
combinations of time and pressure) control to optimize plunger lift performance. Plunger
performance data, such as number of plunger travels, total plunger arrivals, and plunger travel
time, can be stored in the controller. This controller uses very little power, which is supplied by a
small solar power system with available battery backup. Auxiliary inputs, such as pressure
sensors, can also be monitored by this system. The system can be accessed and monitored
remotely using radio, cellular telephone, or satellite links.

SYSTEM MONITORING
One common method of monitoring the performance of a plunger lift well is a surface pressure
recorder. Analyzing the tubing and casing pressure will show whether the plunger is working as it
should.
The operations of the controller should also be monitored. If the controller does not function
properly, fluid and gas recovery will be affected. Modern computer-based systems can monitor the
status and operational parameters of the controller.
The plunger lift monitoring system uses a controller, downhole and surface sensors and
communications system to monitor the downhole operating parameters. The controller, powered
by a solar array with battery backup, can collect operational data from casing and tubing pressure
sensors or other surface sensor systems. The well sensors typically monitor plunger speed,
battery voltage, plunger arrivals, and other operational parameters. The communications system
transmits data, automatically or on demand, to and from the controller. This system can use radio,
landline or cell phone, and satellite.
Monitoring the well lets the field operator identify problems in real time. Using the data sent from
the well, the operator can assess operational problems, and even predict and correct future
problems. The monitoring system also lets the operator remotely adjust such functions as the
plunger lift controller, thus optimizing field performance without going to the field.
Inflow performance of the well should be measured periodically to ensure that the plunger lift
system is operating as it was designed. The plunger itself should be inspected to see that it is not
worn or damaged. If the external sealing elements are damaged, fluid production will decline and
the injected gas will be wasted.
Another important factor in the successful use of a plunger lift system is the skill and experience of
the field operating personnel. Since plunger lift is a mechanical process, field operators must
spend more time with a well on plunger lift. Plunger lift requires careful and frequent monitoring to
reduce or eliminate operating problems.

DETERMINING WELL SUITABILITY


Determining the suitability of a well for plunger lift and what type of flow system should be used
are major problems in plunger lift applications. A method for making this evaluation was defined by
Foss and Gaul. The following information is needed to make such an evaluation:

Depth of the well


Wellhead back pressure
GLR (in MCF/barrel)
MCF required per cycle
Required average surface casing pressure
Maximum production attainable in b/d
Maximum cycles per day.

It is common practice for service companies to assist the engineer in making this determination,
including the proper sizing of downhole and surface equipment.
Another major consideration when evaluating a well or field for plunger lift is the supply of gas for
injection. All the energy needed to lift the plunger and produced fluids comes from gas, so if there
is not enough gas in the formation fluid or available from the field separator system, it must come
from another source. The fluid GLR is measured to determine whether additional gas from
external supply sources will be needed. This is an important consideration when evaluating the
feasibility of installing plunger lift on a well and may make plunger lift an uneconomic option in
some cases.
Other considerations when evaluating a well for possible plunger lift are the diameters of the
master valve and the tubing. They must have the same diameter so that the plunger can move into
the lubricator easily and can fall freely at the start of the next lift cycle. Conditions that can make a
well a poor candidate for plunger lift are tight spots or bends in the tubing, certain types of gas lift
valve mandrels, or sand production. High well deviation also may keep the plunger from moving
freely through the tubing. Always gauge the tubing and check the measurements of the surface
equipment before running the plunger so that any tubing restrictions can be identified.

Das könnte Ihnen auch gefallen