You are on page 1of 84

Completion Equipment

Tubulars
API Specifications for Oilfield Tubulars
The American Petroleum Institute (API) has defined certain standards for oilfield
tubular goods, such as tubing and casing. The API has defined ten grades of steel:
H40, J55, K55, C75, L80, N80, C90, C95, P105, and P110. The number indicates the
API minimum yield strength in thousands of psi. The letters H, J, and N are primarily
to minimize verbal confusion, while the others have an additional meaning:
K has higher ultimate strength than J
C, L "restricted yield strength" with tighter specifications
P high strength
The behavior of tubular goods under stress conditions is a basic problem in strength
of materials. The API has developed a set of standard formulas that are used
throughout the oil industry to predict the minimum load-carrying capacity to be
expected from a particular grade and weight of pipe (API Bulletin 5C3). Tables of
casing and tubing strengths based on the formulas are also published by the API
(Bulletin 5C2) and in various manufacturers' and service companies' handbooks.
Remember that the API formulas are modified from time to time and it is important
to make sure that the performance data used is taken from the most recent version.
The major failure modes that we are concerned with are
burst
collapse
tension failure of the coupling or pipe
There is always some debate as to whether the API formulas are the best theoretical
basis for computing a particular strength parameter (e.g., for burst, a modified
Barlows equation is used instead of Lame). However, each company's assessment of
the conservative nature or inadequacies of the API formulas is generally reflected in
the design factor and design assumptions that they apply in using the API Strength
Criteria.

Tubing Design Concept


The uncertainties regarding actual loading conditions and the state of the tubing
(e.g., corrosion, anomalies due to poor handling) considerably exceed our analytical
capabilities to determine the resultant stresses. The tendency in the oil industry
therefore has been not to be overly sophisticated in analyzing an extremely complex
system, but rather to make designs on the basis of a set of idealized loading
conditions that have proven adequate in the past, such as those presented in Table

1. It is important to remember that each company has its own philosophy, criteria,
and design factors to consider. The balance of design assumptions versus actual
conditions is depicted in Figure 1 (The balance of design assumptions versus actual
conditions).

Figure 1

While this may lead to a tendency to overdesign, the relative cost of the convenience
is generally fairly small. Extreme caution should therefore be used in making
modifications to the idealized loading assumptions. For special, severe loading
conditions (e.g., ultra deep >20,000 ft (6000 m), very high pressure >10,000 psi (70
MPa), very hot >300 F (150 C)), it is necessary to make a detailed computerassisted stress analysis.

Condition

Loading

Design
Criteria

Typical Design Factor

Burst

Internal

Kill pressure on
hydrocarbonfilled tubing

1.125

Collapse

Tension

External

Packer fluid
and zero
annulus
pressure

Considerations

Check effects
of compression

External

Casing head
pressure= shutin tubing
pressure

Internal

Tubing empty
and
depressured

Considerations

Check effects
of tension

Running

Buoyant weight
in completion
fluid

1.125

Body: 1.333

Joint: 1.8*
Tension and
Compression

Operating

Cold
stimulation and
hot production
conditions

Body: 1.125

Joint: 1.333

Total Stress

Considerations

Check effects
of temperature
and pressure
changes

Triaxial

Max. stress

80% yield

*Assuming separate checks are not planned on shock and bending effects; otherwise
use 1.5.
Table 1: Typical criteria for tubing design on a flowing well
In many field situations and preliminary estimates, to establish the weight and
strength of the tubing it is sufficient simply to look at the tubing rating and to apply
the corporate design factor. However, it must be recognized that loading conditions
vary over the length of the tubing string, and to properly visualize this it is generally
advantageous to carry out a graphical tubing string design.
Graphical Tubing String Design

This is a convenient way of understanding loading conditions and presenting design


results. The technique is presented in Example 2 (part 1) and illustrated in Figure 2
(Graphic tubing design estimated operating pressures),

Figure 2

Figure 3 (Graphic tubing design burst loads), and Figure 4 (Graphic tubing design

tubing selection).

Figure 3

Abbreviations are presented in the Nomenclature.

Figure 4

Example 2 (part 1)
Graphical Tubing Design
Planning Data

KBE:

3000 ft (915 m)

TD:

11,500 ft (3500 m)

Tbg:

2 7/8 in. OD (73 mm)

Closed-in bottomhole
pressure:

5500 psi (38 MPa)


estimated from mud weight

Formation breakdown

12,500 psi (86 MPa)

pressure:
estimated from offset well
Fracture propagation
pressure:

9200 psi (63 MPa)


estimated from offset well

Packer fluid:

inhibited oil (0.38 psi/ft)

Production:

expect sour gas


(gas gravity = 0.80 reservoir)
(gas gravity = 0.70 separator)
J55 or L80 tubular to be used

Stimulation:

fracture expected (assume 20 barrels per minute); maximum


allowable annulus pressure is 2000 psi (13,790 kPa)

THP Estimate

Depth of
Hole

Gas Gravity

(ft)

(m)

0.60

0.65

0.70

0.80

1000

305

.979

.978

.976

.973

2000

610

.959

.956

.953

.946

3000

915

.939

.935

.930

.920

4000

1219

.920

.914

.907

.895

5000

1524

.901

.893

.885

.870

6000

1830

.883

.873

.854

.847

7000

2133

.864

.854

.844

.823

8000

2438

.847

.835

.823

.801

9000

2743

.829

.816

.804

.779

10,000

3048

.812

.798

.764

.758

11,000

3353

.795

.780

.766

.737

12,000

3660

.779

.763

.747

.717

13,000

3962

.763

.746

.729

.697

14,000

4267

.747

.729

.712

.678

15,000

4572

.732

.713

.695

.659

16,000

4876

.717

.697

.670

.641

17,000

5181

.702

.682

.652

.624

18,000

5486

.687

.656

.645

.607

19,000

5791

.673

.652

.631

.590

20,000

6097

.659

.637

.615

.574

Table 2: Ratio between surface pressure and bottomhole pressure in gas wells for a
range of gas gravities
At a gas gravity = 0.8, CITHP = 0.727

CIBHP = 3999 psi

At a gas gravity = 0.7, CITHP = 0.757

CIBHP = 4164 psi

For a kill situation:


bottomhole injection pressure = CIBHP + 2000 psi = 5500 psi + 2000 psi = 7500
psi
If gas gravity = 0.8, THIP = 0.727

BHIP = (0.727) (7500)

=5453 psi
Assumed Fracture Conditions
1. Formation breakdown achieved with water
2. Fracture job carried out with water-base fluid
Friction loss in 2 7/8 in. tubing at 20 BPM using water with friction reducer is 350 psi/1000 ft for
11,500 ft (Dowell Handbook)
FPP = 9200 psi
Friction = +4025 psi (350
Head = -5175psi (0.45

11.5)
11,500)

Frac THP= 8050 psi


Prepare a depth pressure plot ( Figure 2 ) in the following manner:
1. Plot the closed-in bottomhole pressure (CIBHP).
2. Plot the formation breakdown pressure (FBP) and the fracture
propagation pressure (FPP).

3. Plot the packer fluid gradient, fracture fluid gradient, and water
gradient.
4. Estimate wet and dry gas gradients and plot these up from the
closed-in bottomhole pressure.
5. Establish the closed-in tubing head pressure for normal production
conditions (i.e., oil or, as in this case, wet gas) and for worst case
design assumption (usually dry gas).
6. Establish maximum THP for which completion is to be designed,
which normally will be kill or stimulation conditions (fluid gradient
through FBP, FPP, or specified differential above CITHP). For Example
2, the graphical design should now look like Figure 2 .
7. Establish through inspection the greatest differential pressure at
surface and downhole (usually stimulation conditions). Determine what
steps can be taken to reduce loading (e.g., maintaining maximum
allowable annulus pressure during stimulation). Plot adjusted annulus
pressure line ( Figure 3 ).
8. Plot burst load line (BLL) as difference between most critical tubing
and annulus pressures. The BLL is a function of the relative densities in
the tubing and annulus. BLL will generally, but not always, decrease
with depth ( Figure 3 ).
9. Plot critical collapse load conditions (CLL). Normally we assume that
a slow leak has changed the CHP to CITHP and that tubing is empty
and depressured. This can occur in gas wells if the tubing becomes
plugged or a downhole safety valve is closed. Conditions can approach
this situation in oil wells after a fracture treatment if operators
commence kickoff before bleeding off annulus pressure. (In some
cases this may be a more critical load ( Figure 4 ).)
10. Plot pressure test conditions (PT). This is often the most critical
load to which a completion is subjected. Consider timing of the
pressure test and density of fluids in the tubing and annulus at time of
test.
11. Look up tubing performance data in API Bulletin 5C2.
12. Adjust API internal yield (burst) and collapse resistance
specifications with design factor (see Figure 1 and API Bulletin 5C2).
13. List resulting tubing capabilities ( Figure 4 ).
14. Compare design loads with tubing capabilities and select tubing. In
most cases the optimum tubing grade and weight will vary with depth.
To minimize costs and/or tensional loads, such variations may be
incorporated, although there will then be a constraint on pressuretesting capabilities. However, most operators prefer to use a common

weight and grade throughout the completion, if possible. This reduces


the risk of installation and operating errors. When regulations permit,
the designer may be able to compromise slightly on accommodating
loading conditions deep in the hole, if the associated design
assumption is extremely unrealistic (e.g., a completely empty tubing in
a high productivity oil well). However, the designer must first check on
how critical the actual biaxial (or triaxial) loading conditions are likely
to be and make appropriate notes in the well file.
With reference to Example 2, in Figure 4 the options include the following:
1. full string of 6.4 lb/ft L80 tubing
2. 0 to 6500 ft = 6.4 lb/ft J55
6500 to TD = 6.4 lb/ft L80
3. full string of 6.4 lb/ft J55 with modified collapse design criteria of
2000 psi as maximum CHP with an empty tubing
Since 2000 psi is the maximum allowable annulus pressure during stimulation, option 3 may be
an acceptable design. Since the differential cost of J55 and L80 is around $3 per ft, the potential
saving of $34,500 between options 1 and 3 may justify further detailed engineering work. On the
other hand, if the wellstream is expected to be extremely corrosive, the higher grade tubing may
be selected in any case to provide a corrosion allowance.
The key things to note from Figure 5 (Effect of buoyancy on axial load) are
the most severe burst loadings occur at surface

Figure 5

the most severe burst and collapse loadings occur during pressure testing,
well kill, and stimulation
the most severe collapse loading occurs downhole
additional annulus pressure can be used to reduce burst loading, provided
the casing is strong enough
the tubing-head pressure during kill operations (THIP) often approximates or
exceeds the reservoir pressure (CIBHP)
With relatively small tubing strings (<3.5 in. or 90 mm), the inherent burst and collapse strength is
so high that some engineers do not bother with tubing design in wells with depths of less than
8000 ft (2500 m), unless overpressures are expected.
Simplified Tensional Strength Design
Although burst and collapse resistance may not be significant considerations in
pumping wells, tensional strength is a critical design parameter for all wells. Coupling
leakage and failure, which accounts for 80% of the problems in well tubulars, often

may be the result of inadequate tensional design rather than a burst or sealing
problem. In this respect, it is particularly important to remember that test pressures
impose substantial piston forces on the tubing (e.g., a 2000 psi (13.8 MPa) pressure
test on a plug set inside 2 7/8-in. (73-mm) tubing will increase the tension on the
hanger by (2000)( /4)(2.44l)2 = 9360 lb (42 kN).
It is also important to recognize that, unlike other strength parameters, the API joint
strength is based on a failure condition rather than the onset of plastic deformation.
The failure condition is either an unzipping of the pin and box in the case of API
threads, because of yielding (also called "jump-out"); or breakage of reduced cross
section at the threads in the case of square threads.
Finally, there are all sorts of additional tensional loads that we do not normally
analyze in detail (e.g., shock loading and drag forces during running, bending
stresses, buckling, cross-sectional piston forces, changes in buoyancy).
Since it is common practice to make a preliminary tensional design using tubing
weight loading only, a higher design factor is used for tension and especially for joint
strength (Table 1, above). Some companies and more conservative engineers will
even ignore the potential benefits of buoyancy. Buoyancy results in a piston force on
the lower end of the tubing and as a first approximation it may normally be assumed
that

(12)
where:
WB = buoyant weight
WN = weight in air
= density of steel ( 8 gm/cc)
= density of fluid
Figure 5 graphically depicts the tensional and compressional forces at work on a tapered string of
tubular goods. The load resulting from the weight of the pipe is shown for a string weighed in air,
and with the buoyant forces accounted for as piston forces or approximated using Equation 12.
We can see that at a point approximately midway in the length of the heavier pipe at the bottom
of the string there is a change from compression to tension. This is also the concept which guides
the design of drillstrings with the purpose of keeping the drillpipe in tension while using the
heavier drill collars to maintain a compressional load on the bit.
Part 2 of Example 2 gives the preliminary tension design considerations for the
completion already covered in part 1.
Example 2 (part 2)
Preliminary Tension Design

Tubing weight: 6.4 lb/ft


Tubing length: 11,500 ft
Packer fluid: inhibited oil
0.38 psi/ft = 0.88 gm/cc
WN = 6.4 11,500
= 73,600 lb

= 0.89

73,600

= 65,504 lb
Joint Specifications
J55

L80

EUE

HYD CS

EUE

HYD A95

API joint strength


(Klb)

99.7

100

135.9

128

Design factor
(Table 1)

1.8

1.8

1.8

1.8

Design capacity
(Klb)

55.4

55.6

75.5

71.1

Tubing Tension Design Considerations


1. Requires L80 tubing at surface
2. Requires joint strength capability of EUE or equivalent
3. In view of pressures, depth, and H2S would probably select premium
grade coupling
Many companies have these design techniques programmed for the computer and use the same
general technique for both tubing and casing designs.
Tubing Design Parameters
It is important to remember that while the primary function of the tubing is as a
conduit for hydrocarbon production or for injection of water or gas, the most severe
loadings often occur during well service or killing operations, or during pressure
tests. It is therefore prudent to make provision for these operations when designing
a completion, and to check out the tubing limitations when planning a well servicing
operation (e.g., a stimulation or a workover) . Care must be taken not to increase
completion costs excessively by trying to make provisions for all sorts of unlikely, but
possible, occurrences. It must also be remembered that there are steps that can be

taken to mitigate the induced stresses during many operations (e.g., applying
annular pressure or heating fracturing fluids). On the other hand, the consequential
costs of a failed tubing string, or of having to run a special working string, in terms
of deferred production and rig time, can be quite substantial. Assessment of the
most cost effective solution is generally a judgment call based on the engineer's
experience and on corporate attitudes and policy. A typical set of parameters has
already been illustrated in Example 2.
Burst
The tubing and wellhead should be designed for squeeze and kill conditions. Since
fines in the perforations or oil can sometimes cause a "check valve" effect when
attempting to squeeze back liquids, many completion designers like to have the
flexibility of being able to raise the bottomhole pressure to the FBP or at least to the
FPP. However, with high permeability reservoirs or gas wells in which fracture
stimulation is unlikely, completion engineers are often satisfied with a certain
minimum differential for injection. The value selected varies from area to area and
from company to company, but is commonly either around 1000 psi (7 MPa), or 33%
of the reservoir pressure. The author suggests
1. FBP where k1 < 100 md
kg < 50 md
2. FPP for squeezing liquids, where k1 > 100 md
3. CIBHP + 1000 psi (7 MPa) for squeezing gas, where
kg > 50 md; or for squeezing liquids, where k1 > 1000 md
From the rock mechanics theory presented by Geertsma (1978) and others it may be deduced
that in a tectonically relaxed area, a provisional estimate of the fracture propagation gradient
(FPG) can be obtained from the equation

(13)
FPG < FBG < 1.1 psi/ft (25 kPa/m) (14)
where:
sv = overburden stress ( ~1 psi/ft depth)
p = pore pressure, psi
D = depth, ft
FPG = formation propagation gradient, psi/ft
FBG = formation breakdown gradient, psi/ft
The specification of the pressure test conditions is often critical to burst design. Government
regulations sometimes specify pressure test conditions (e.g., to at least 90% of the reservoir
pressure or to 1000 psi (5 MPa) over the maximum differential pressure expected at the packer).
If no regulations exist, most operators test to their tubing design conditions.

Collapse
Severe collapse loads on the tubing can occur
in gas wells and high GOR oil wells with low-flowing bottom-hole pressures and deepset safety valves, after blowdown to test a plug, etc.
during annulus pressure tests, or operation of shear circulation devices
where there are pressured annuli
during underbalance perforating or testing at high drawdown
during tubing blowouts
It is important to remember that tension reduces collapse strength. This biaxial effect should be
examined for large diameter tubings, especially if reduced collapse design assumptions and/or a
deep-set safety valve is used.
Tension
Tubing strings are not only subjected to running tensions with all the associated
shock and acceleration loadings, but also to varying operating stresses due to piston
forces on the steel and/or any plugs, pumps, standing valves, and the like in the
tubing. Moreover, if the tubing is anchored or held by a packer, its operating tension
will vary as a result of
thermal effects (hot production or cold kill fluid)
piston effects (changes in buoyancy and forces at joint upsets)
ballooning effects (changes in internal or external pressure)
buckling effects (longitudinal instability)
These potential problems are listed in terms of their most common relative magnitude (although
the relative importance of piston and ballooning effects is variable).
Combined Loading
While the designer of tubular goods normally talks in terms of burst, collapse, and
tension compression as if they were independent, it is obvious that in most actual
loading situations they occur simultaneously. Precise stress analysis should really
consider a triaxial loading situation.
The simultaneous solution of all the associated equations is rather complicated. A
number of computer programs are available, but for most field engineers they will be
a "black box" solution. This can be dangerous. It is important to check that the
formulas are properly handled, particularly with respect to collapse, which is a
"stability effect." Therefore it is usual for critical stress analyses (e.g., for ultra deep,

high pressure, or sour wells) to be undertaken by a specialist consultant, research


group, or intracompany task force. Moreover, since this is not the routine design
technique, design factors are less well proven (although a value of 1.25 is often
used).
A more convenient approach for the intermediate range, moderately complex design
problem is to use the ellipse of biaxial yield stress proposed by Holmquist and Nadai
(1939). The critical relationships are (a) tension reduces collapse resistance; (b)
compression reduces burst resistance.
The other important concept in the consideration of triaxial loads is that pressure
changes affect axial stresses or cause tubing movement. This has been extensively
discussed in SPE papers by Lubinski (1962), Hammerlindl (1977), and Stillebroer
(1967).
Bending
Bending stresses can be significant in large tubulars. They are compressive in the
inner wall and tensional in the outer wall, the most detrimental being

(15)
where:
R = the radius of curvature (ft)
sb = bending stress
E = Young's modulus (for steel, E = 30

106 psi)

do = outside diameter of the tubular


Bending stresses result from both hole curvature and buckling. The effects of doglegs need only
be considered if they are very severe (>10/100 ft; 10/30 m) or if very large tubing (5 1/2 to 7 in.;
140 to 178 mm) is being used.

Production Casing
The production casing must be adequately sized for the planned completion. It will
obviously affect the size of the other required casing strings, the bit selection, the
capacity of the rig, and the overall well costs. The production casing must be
designed for the loads that may be imposed during the producing life of the field. It
is similar to tubing design in several ways.
Burst
Production casing must be designed to withstand the maximum closed-in tubing
pressure that can be expected. If a packer has been used, this pressure is assumed
to be applied at the top of a full column of packer fluid (i.e., for the case of a tubing

failure at the surface). We usually assume that the external pressure resisting burst
is a water gradient. If the packer fluid is heavier than water the burst load will
increase with depth.
In the event that a snubbing operation could not be conveniently attempted if a
tubing break occurs at the surface, the casing must be strong enough to withstand a
bullhead squeeze on the live tubing string, in which case this would be the design
criteria for the casing and wellhead.
In many cases it may be necessary to design the casing for loads imposed during
stimulation and pressure testing. Conversely, the casing capacity must be checked
when designing a fracturing treatment. This is particularly important in wells where
no packer is used.
Collapse
Production casing may be subject to complete evacuation during production
operations if the well is operated on gas lift or pumped-off, or if the packer or
workover fluid is lost into a depleted zone. As the pipe may have deteriorated before
this occurs, a higher design factor (1.125+) is often used for production casing.
Severe collapse loads may occur in situations in which thermal expansion of the
annular fluid between the production and intermediate strings cannot be bled off
(e.g., in some subsea wells).
Increased loading should be assumed if live annuli are a feature of the area. Reduced
loadings may be assumed if the wells will not be pumped off, gas lifted, or severely
depleted.
Severe collapse loads may exist in the pay section during high drawdown,
underbalanced perforating and testing, and squeeze either or both cementation and
fracturing ( Figure 1 , Collapse loads in the pay). It is highly advisable to maintain
some set casing/tubing annulus pressure during such service operations.

Figure 1

Tension/Compression
In high rate production areas and thermal wells, expansion of the production tubing
may impose additional tension on the casing strings, via the packer.
Couplings
In high pressure (>5000 psi; 34 MPa), high temperature (>300 F; 422 K) and/or
severely sour conditions, premium casing couplings are recommended.
Material Selection
In sour environments, material specification must consider the chances of H2S
contamination of the casing/tubing annulus and the added possibility of temperature
changes during stimulation affecting the stress corrosion tolerance of the pipe.
Couplings
There are many forms of coupling available, some of which have been dedicated to
the public through the auspices of the API, while others are produced by, or under

license from, a specific manufacturer. It was, in fact, the need to obtain a


standardization of thread forms and diameters that led to the formation of the API
Committee on Standardization of Tubular Goods in 1924.
The API couplings ( Figure 1 ,

Figure 1

Figure 2 and Figure 3 ,

Figure 2

Cutaways of basic types of couplings) are of three basic types: external upset (EUE),
nonupset (NU), and integral joint.

Figure 3

Threads are of two main forms. The round API threads are weaker than the pipe
body (i.e., <100% efficient). The buttress threads were developed by the National
Tube Division of the United States Steel Corporation to provide a high strength
coupling for deep, high pressure well service. This thread is used in a number of
proprietary couplings, e.g., Hydril, VAM, Atlas-Bradford.
All tapered threads achieve a seal by driving the pin and box surfaces together under
sufficient stress to generate a bearing pressure exceeding any differential that is to
be subsequently applied. However, a small spiral void is always left between the
mating surfaces, and must be filled with solids in the form of thread compound. (This
is the reason for careful specification of the compound in Bul 5 A2.) On API round
threads, this void occurs between the crest and root of the mating threads, while on
buttress threads it extends over the whole flank of the thread on the beveled side.
To improve leak resistance, especially at elevated temperatures and under high
pressure differential, the so-called "premium" seals were developed. These consist of
either metal-to-metal seals on tapered portions of the pin and box surfaces or an
elastomer seal ring, or both. This type of seal requires a high quality finish and
precise gauging and inspection. The coupling is therefore more costly. Since API and

buttress threads have proven to be very reliable in the field, the decision to use the
more expensive "premium" seals requires careful economic justification. In general,
their application has proven valuable in highly corrosive conditions in high pressure
gas wells or high pressure/high GOR oil wells, and in thermal wells subject to high
compressive loads. They may be used where workover costs are high (e.g.,
offshore). In general, API tubulars are adequate for differential pressures of less than
5000 psi (34.4 MPa) and temperatures of less than 300 F (150 C), using high
temperature thread compound. For corrosive conditions and continuous gas service,
the pressure limit is often reduced to 2500 psi (17 MPa).
Exercises

Oilfield Units
The 2 7/8-in., 6.4 lb/ft, J-55 NU tubing string in a 6000-ft oil well was designed
based on the tubing's buoyant weight in water. What additional load would be
imposed on the tubing if a rod pump were to be pressure tested to 500 psi after the
well had been operating for some time and the annulus pumped off? The well
produces 30 API oil (SG: 0.876) with no water.
If company policy dictates the use of a design factor of 1.5 for joint strength, can
this pressure test be safely carried out?

SI Units
The 73-mm (2 7/8-in.), 9.5 kgm/m (6.4 lb/ft), J-55 NU tubing string in an 1830-m
oil well was designed based on the tubing's buoyant weight in water. What additional
load is imposed on the tubing if a rod pump were to be pressure tested to 3.5 MPa
after the well had been operating for some time and the annulus pumped off? The
well produces 876 kg/m3 oil (SG 0.876) with no water.
If company policy dictates the use of a design factor of 1.5 for joint strength, can
this pressure test be safely carried out?
Solutions

Oilfield Units
Weight in air = 6000 ft x 6.4 = 38,400 lb

Buoyancy in water =

= 0.875

(density of water 1 gm/cc; density of steel = 8 gm/cc; use Equation 12)

(12)

Buoyant weight when run 38,400 x 0.875 = 33,600 lb


Water gradient (SG = 1) = 0.433 psi/ft
Oil gradient (SG = 0.876) = 0.876 x 0.433 = 0.379 psi/ft
ID of tubing = 2.441 in., Ai = 4.680 in2
Weight of oil in tubing = 6000 x 0.379 x 4.680 = 10,642 lb
Total applied tensile force at the surface, after well has been pumped off, equals
Weight of oil + weight of tubing in dry casing
+ applied pressure x area of tubing ID
= 10,642 + 38,400 + 500 (4.680)
= 10,642 + 38,400 + 2340
= 51,382 lb
The API joint strength rating of 2 7/8-in., 6.4 lb/ft, J-55, non-upset tubing = 72,600
lb.
If a design factor of 1.5 for joint strength is dictated by company policy, the
allowable tensile loading (with design factor) = (72,600) / (1.5)
= 48,400 lb.
Therefore, the proposed loading exceeds the design specifications and a leak might
occur during the test. In fact, the loading exceeds design specs even without adding
the 500 psi of wellhead pressure (actual load 49,042 lb compared with an allowable
of 48,400 lb). However, using a 1.33 design factor for operating loading conditions,
pressure tests of up to 1184 psi would be permissible:
(72,600) (1.33) = 10,642 + 38,400 + pt (4.68)
Pt = 1184 psi

SI Units
Weight in air = 1830 x 9.5 = 17,385 kg = 170.5 kN

Buoyancy factor in water =

= 0.875

(density of water = 1 gm/cc; density of steel = 8 gm/cc; use Equation 12)

(12)
Buoyant weight when run = 170.5 x 0.875 = 149.2 kN
Water gradient (SG = 1.0) = 9.794 kPa/m
Oil gradient (SG = 0.876) 0.876 x 9.794 = 8.579 kPa/m

ID of tubing = 62 mm, thus Ai = 30.19 x l0-4 m2


Weight of oil in tubing = (1830 m) (8.579 kPa/m) (3.019 x 10-3 m2)
= 47.4 kN
Total applied tensile force at the surface after well had been pumped off equals
Weight of oil + weight of tubing in dry casing
+ applied pressure x area of tubing ID
= 47.4 + 170.5 + 3500 (3.019 x 10-3)
= 47.4 + 170.5 + 10.57
= 228.47 kN
From Figure 1 , the API joint strength rating of 73-mm, 9.5 kgm/m, J-55, nonupset tubing is
72,600 lbs (322.92 kN).

Figure 1

If a design factor of 1.5 for joint strength is dictated by company policy, the
allowable tensile loading (with design factor)
= 322.92/1.5
= 215.28 kN

Therefore, the proposed loading exceeds the design specifications and a leak might occur during
the test. In fact, the loading exceeds design specifications even at zero wellhead pressure (actual
load of 47.4 + 170.5
= 217.90 kN compared with an allowable load of 215.28 kN).
However, using the 1.33 design factor for operating loading conditions, pressure tests
of up to 8252 kPa would be permissible:

= 217.90 + pt (30.19 x l0-4)


pt = 8248 kPa

2.. Oilfield Units


A 12,000-ft, hydrostatically pressured gas well ( g= 0.7) is to be completed. What closed-in tubing
head pressure can be expected? If the casing, wellhead, and tubing are to be designed for a
squeeze kill at fracture propagation pressure (FPP) what wellhead pressure would be expected at
the start of the kill operation? (Use Equation 13 for estimating the FPG.)

(13)

SI units
A 3660-m hydrostatically pressured gas well ( g = 0.7) is to be completed. What closed-in tubing
head pressure is expected?
If the casing, wellhead, and tubing are to be designed for a squeeze kill at fracture
propagation pressure (FPP) what well-head pressure would be expected at the start
of the kill operation? (Use Equation 13 for estimating the FPG.)

Solutions

Oilfield Units
Condition

Loading

Design Criteria

Typical Design Factor

Burst

Internal

Kill pressure on
hydrocarbon-filled tubing

1.125

External

Packer fluid and zero


annulus pressure

Considerations

Check effects of
compression

External

Casing head pressure=

Collapse

1.125

shut-in tubing pressure

Tension

Tension and
Compression

Internal

Tubing empty and


depressured

Considerations

Check effects of tension

Running

Buoyant weight in

Body: 1.333

completion fluid

Joint: 1.8*

Cold stimulation and hot


production conditions

Body: 1.125

Operating

Joint: 1.333
Considerations

Total Stress

Check effects of
temperature and
pressure changes

Triaxial

Max. stress 80% yield

*Assuming separate checks are not planned on shock and bending effects; otherwise
use 1.5.
Table 1: Typical criteria for tubing design on a flowing well
The ratio of THP to BHP for

of 0.7 = 0.747.

At 12,000 ft, assuming a hydrostatic gradient of 0.433 psi/ft, the bottomhole


pressure can be estimated as
BHP = 12,000 x 0.433 = 5196 psi
with gas surface pressure 5196 x 0.747 CITHP = 3881 psi
Assuming the overburden gradient to be 1 psi/ft and using Equation 13 to estimate
the fracture propagation pressure:

= 0.717 psi/ft
FPP = 0.717

12,000 ft = 8604 psi

Maximum wellhead pressure at the start of the kill operation (THIP) will equal the formation
propagation pressure minus the pressure due to the gas gradient:

THIP = (12,000
0.717) - (5196 - 3881) = 7289 psig
More correctly, we should use the ratio in Table 1 since the increased pressure will increase the
gas density.
THIP = (12,000
0.717)
0.747 = 6427 psi
Therefore, a 5000 psi wellhead should not be used even though the maximum closed-in tubing
head pressure is only about 4000 psi.

SI Units
Referring to Table 1:
The ratio of THP to BHP for

= 0.7 = 0.747

At 3660 m, assuming a normal hydrostatic pressure of 9.795 kPa/ m, bottomhole


pressure can be estimated as
BHP = 3660
9.795 = 35,850 kPa
with gas surface pressure = 35,850
0.747
CITHP = 26,780 kPa
Assuming the overburden gradient to be 22.62 kPa/m and using Equation 13 to
estimate the fracture propagation pressure:

= 16.21 kPa/m
FPP = 16.21

3660 = 59,329 kPa

Maximum wellhead pressure at the start of the kill operation


(THIP) will equal the formation propagation pressure minus the pressure due to
the gas gradient:
THIP = (3660

16.21) - (35,850 - 26,780) = 50,259 kPa

More correctly, we should use the ratio in Table 1 since the increased pressure will increase the
gas density:
THIP = (3660
16.21)
0.747 = 44,319 kPa
A 64 MPa wellhead should therefore be used.

3Oilfield Units
A 7000-ft well that is to be produced with a target of 15,000 STB/d using 5.5-in.
tubing encounters 170 ft of oil-bearing formation with a pressure of 3000 psi. What
rating of wellhead should be used? If a single grade and weight of tubing is to be

used, what is the cheapest string that can probably be run, assuming that

Grade

Weight
(lb/ft)

Collapse
Strength
(psi)

Burst
Strength
(psi)

Tensional
Strength
(1000 lb)

Cost Comparison

J-55

15.5

4040

4810

300

cheapest

17.0

4910

5320

329

C-75

17.0

6070

7250 423

N-8O

17.0

6280

7740

446

20.0

8830

8990

524

most expensive
moderately expensive

packer fluid: inhibited seawater (gradient = 0.435 psi/ft)


reservoir pressure: 3000 psi
kill pressure: > 1000 psi above CIBHP (squeezing gas)
< fracture propagation pressure (squeezing oil)
Fracture propagation pressure gradient (FPG) is approximately:

FPG = 0.5 + 0.5

(psi/ft)

If gas cap gas should break through into the well, assume gas
gravity = 0.65.
Use Table 1 to calculate head of gas.

Condition

Loading

Design Criteria

Typical Design Factor

Burst

Internal

Kill pressure on
hydrocarbon- filled
tubing

1.125

External

Packer fluid and zero


annulus pressure

Considerations

Check effects of
compression

External

Casing head pressure=


shut-in tubing pressure

Internal

Tubing empty and

Collapse

1.125

depressured
Considerations

Check effects of tension

Tension

Running

Buoyant weight in
completion fluid

Body: 1.333
Joint: 1.8*

Tension and
Compression

Operating

Cold stimulation and hot


production conditions

Body: 1.125
Joint: 1.333

Considerations

Check effects of
temperature and
pressure changes

Total Stress

Triaxial

Max. stress 80% yield

*Assuming separate checks are not planned on shock and bending effects; otherwise
use 1.5.
Table 1: Typical criteria for tubing design on a flowing well.
Assume live oil gradient is 0.3 psi/ft.
Consider only the maximum burst at the wellhead, the maximum collapse load at the
packer, and the running tension.

SI Units
A 2130-m well that is to be produced with a target of 2400 m3/day of oil with l40mm (5 1/2-inch) tubing encounters 50 m of oil-bearing formation with a pressure of
20,700 kPa. What rating of wellhead should be used? If a single grade and weight of
tubing is to be used, what is the cheapest string that can probably be run, assuming
that

Grade

Collapse
Strength

Burst
Strength

Tensional
Strength

(kPa)

(kPa)

(kN)

23.1

27,855

33,164

1,334

25.3

33,853

36,680

1,463

C-75

25.3

41,851

49,987

1,882

most expensive

N-80

25.3

43,299

53,365

1,984

moderately expensive

29.8

60,880

61,984

2,331

Weight
(kgm/m)

J-55

Cost Comparison

cheapest

packer fluid: inhibited seawater (gradient = 9.840 kPa/m)


reservoir pressure: 20,700 kPa
kill pressure: > 7,000 kPa above CIBHP (squeezing gas)
< fracture propagation pressure (squeezing oil)
Fracture propagation pressure gradient (FPG) is approximately

FPG = (0.5) (22.3) + 0.5


kPa/m
If gas cap gas should break through into the well, assume gas
gravity = 0.65.
Use Table 1 to calculate head of gas.
Assume live oil gradient is 6.786 kPa/m.
Consider only the maximum burst at the wellhead, the maximum collapse load at the
packer, and the running tensions.

Solution

Oilfield Units
CITHP = CIBHP -

o = 0.3 psi/ft
under operating conditions:
CITHP = 3000 - (0.3 7000)
= 900 psi
under maximum conditions (assuming gas breakthrough)
CITHP = 0.854 x BHP (from Table 1)
MAX CITHP = 0.854 x 3000 = 2562 psi
under kill conditions:

FPG = 0.5 + 0.5


= 0.5 + 0.5 x
= 0.71 psi/ft
FPP = 7000 0.71
= 5000 psi

MAX BHIP
(oil kill)
= 5000 psi
equivalent THP
(oil kill)
= 5000 - 0.3 x 7000
= 2900 psi
In the case of a gas kill:
MIN BHIP
(gas kill)
= 3000 + 1000 psi
= 4000 psi
equivalent THP
(gas kill)
= 0.854 x BHP
= 0.854 x 4000
= 3416 psi
From this we see that, while we could probably get away with a 3000 psi wellhead, strictly
speaking we should be using a 5000 psi wellhead.
Incorporation of the provision, in some regulations, for a wellhead rating equivalent
to the reservoir pressure reflects this typical design for kill capability.
Tubing Burst Rating
burst rating = max THP

design factor

design factor = 1.125


required burst rating (for gas kill conditions) = 3416

1.125
= 3843 psi

All grades and weights satisfactory.


Tubing Collapse Loading Calculation
Calculate first assuming a high risk of gas breakthrough:
head of packer fluid = ( f) (D)
= (0.435) (7000)
= 3045 psi
max CHP = max CITHP (assumes leak)
= 2562 psi
max BH annulus pressure = 2562 + 3045
= 5607 psi

max collapse load = 5607 - 0 psi (assumes empty tubing)


= 5607 psi
design factor = 1.125
required collapse rating = 5607 x 1.125 psi
= 6308 psi
Therefore, strictly speaking, we should use 20-lb/ft, N-80 tubing, although it is more likely a l7.0lb/ft would be selected since the probability is very low that such severe collapse-loading
assumptions would prove true.
Calculate next assuming a low risk of gas breakthrough:
max CHP = operating CITHP
operating CITHP = 900 psi
max BH annulus pressure = 900 + 3045 psi
= 3945 psi
max collapse load = 3945 psi
required collapse rating = 3945 1.125
= 4438 psi
We could select 17 lb/ft, J-55 tubing.
Running Tension Calculations
weight in air = W D
= (17 lb/ft) (7000 ft)
= 119,000 lb
buoyancy factor =
=
= 0.8725
buoyant weight of tubing = 0.8725 119,000 lb
= 104,000 lb
design factor = 1.8
required tension rating = (104,000) (1.8) lb
= 187,000 lb
This can be easily carried by any of the tubing grades listed.

It is typical for collapse design to be the critical factor in high-rate, large-tubing


completions, especially at relatively shallow depths.
Conclusion:
5 1/2-in., 17.0-lb/ft, N-80 tubing meets all criteria for this type of production. Risk of
gas breakthrough must be accurately assessed before choosing between N-80 and J55 grades.
SI Units
CITHP = CIBHP - o

xD

o =6.786 kPa/m
under operating conditions:
CITHP = 20,700 - (6.786 x 2130)
= 6246 kPa
under maximum conditions (gas breakthrough)
CITHP = 0.854 x BHP (from Table 1)
MAX CITHP = 0.854 x 20,700 = 17,678 kPa
under kill conditions:

FPG = (0.5) (22.3) + 0.5


= (0.5) (22.3) + 0.5
= 16.00 kPa/m
FPP = (2130) (16.00)
= 34,080 kPa
MAX BHIP (oil kill) = 34,080 kPa
equivalent THP (oil kill) = 34,080 - (6.786 x 2130)
= 19,626 kPa
In the case of a gas kill:
MIN BHIP
(gas kill) = 20,700 + 7000 kPa
= 27,700 kPa
equivalent THP
(gas kill) = 0.854 x BHP
= 0.854 x 27,700
= 23,656 kPa
From this we see that, while we could probably get away with a 20.7 MPa wellhead, strictly
speaking we should be using a 34.5 MPa wellhead.

Incorporation of the provision, in some regulations, for a wellhead rating equivalent


to the reservoir pressure reflects this typical design for kill capability.
Tubing Burst Rating Calculation
burst rating = max THP x design factor
design factor = 1.125
required burst rating (for gas kill conditions) = 23,656 x 1.125
= 26,613 kPa
All grades and weights satisfactory.
Tubing Collapse Loading
Calculate first assuming high risk of gas breakthrough:
head of packer fluid = (f) (D)
= (9.840) (2130)
= 20,959 kPa
max CHP = max CITHP (assumes leak)
= 17,678 kPa
max BH annulus pressure = 17,678 + 20,959
= 38,637 kPa
max collapse load = 38,637 - 0 kPa (assumes empty tubing)
= 38,637 kPa
design factor = 1.125
required collapse rating = 38,637 x 1.125
= 43,467 kPa
Therefore, strictly speaking, we should use a 29.8-kg/m, N-80 tubing, but more
likely 25.3-kg/m tubing would be selected since the probability of such severe
collapse loading is very low.
Calculate next assuming low risk of gas breakthrough:
max CHP = operating CITHP
operating CITHP = 6246 kPa
max BH annulus pressure = 6246 + 20,959
= 27,205 kPa
max collapse load = 27,205 kPa

required collapse rating = 27,205 x 1.125


= 30,606 kPa
We could select 25.3-kg/m, J-55 tubing.
Running Tension Calculation
weight in air = 25.3 2130
= 53,889 kg m
= 528.5 kN

buoyancy factor =
=
=0.8725
buoyant weight of tubing = 0.8725 528.5 kN
= 461.1 kN
design factor = 1.8
required tension rating = (461.1) (1.8)
= 830 kN
This can be easily carried by any of the tubing grades listed.
It is typical for collapse design to be the critical factor in high-rate, large-tubing
completions, especially at relatively shallow depths.
Conclusion:
140-mm (5 1/2-inch), 25.3-kg/m, M-80 tubing meets all criteria for this type of
production. Risk of gas breakthrough must be accurately assessed before choosing
between M-80 and J-55 grades.

Packers
Packer Functions
A packer is a subsurface tool that provides a seal between the tubing and
casing,thereby preventing the vertical movement of fluids across this sealing point.
Packers are used for the following reasons:
to improve safety by providing a barrier to flow through the annulus

to keep well fluids and pressures isolated from the casing


to improve flow conditions and prevent heading
to separate zones in the same wellbore
to place kill fluids or treating fluids in the casing annulus
to pack off perforations rather than use squeeze cementing
to keep gas lift or hydraulic power fluid injection pressure isolated from the
formation
to anchor the tubing
to install a casing pump
to minimize heat losses by allowing the use of an empty annulus or thermal
insulator
to isolate a casing leak or leaking liner lap
to facilitate temporary well service operations (e.g., stimulations, squeezes)
Packer Types
There are many packer manufacturers, some of whom offer an extensive variety of
packers, with each differing to some degree from those of the other manufacturers.
This rather bewildering array can, however, be grouped into principal classes or
types, and may be further categorized by method of setting, by direction of pressure
across the packer, and by the number of bores through the packer.
Packers can be primarily classified as either retrievable, or permanent, or
permanent-retrievable, or inflatable.
Retrievable Packers
This type of packer is run on the tubing ( Figure 1 , Retrievable packer).

Figure 1

After setting, it can be released and recovered from the well on the tubing. Since it is
an integral part of the tubing string, the tubing cannot be removed from the well
without pulling the packer, unless a detachable packer head is used.
Retrievable packers may be designed to be set mechanically or hydraulically.
Mechanical setting methods include rotation of the tubing string, reciprocation of the
tubing string, or the application of tension or set-down weight. With mechanical
packers, the tubing is usually set in compression.
Hydraulic packers are set by applying hydraulic pressure through the tubing string,
but once set they hold the set position mechanically. The tubing is usually in tension.
Retrievable packers are usually used for complex multizone and multistring
completions. Their main limitation was in their limited ability to accommodate tubing
stress changes without unsetting; the availability of effective slip joints and
detachable heads has eased this situation. An historical problem was failure of the
internal elastomer seals, but this technology also has improved markedly in the last
decade. All metal-to-metal seal packers are available but are expensive.

One disadvantage of retrievable packers is that if they fail to retrieve, they must be
removed by milling them out of the casing with an abrasive milling head and a
drillstring. They are very difficult to mill. Generally, this type of packer is used under
nonsevere conditions (differential pressures less than 5000 psi (34.4 MPa),
temperatures less than 300 F (422 K).
Because of the setting mechanism, retrievable packers tend to have a restricted
bore, compared with other packers designed for the same casing size. This factor
may restrict flow or limit wireline operations below the packer depth.
Permanent Packers
Permanent packers are independent of the tubing and may be run on tubing or on
wireline ( Figure 2 and Figure 3 ).

Figure 2

The tubing can be released from the packer and can be pulled, leaving the packer set
in the casing.

Figure 3

Tubing can subsequently be run back and resealed in the packer. The packer may
thus be considered an integral part of the casing. It is sometimes called either a
production packer or a retainer-production packer.
The permanent packer cannot be recovered as such, but it can be destructively
removed (e.g., by milling) ( Figure 4 and Figure 5 ).

Figure 4

If the packer includes a tailpipe and must be recovered, a millout extension is


needed on the packer for the "packer picker," or catch sleeve, on the mill to engage.
In other cases, it may be adequate to simply push the packer to the bottom of the
casing after milling.

Figure 5

Permanent packers can be set using an electric wireline setting tool, a hydraulic
setting tool run on drillpipe or tubing, or by a combination of rotation and pull.
Permanent packers are typically used when
formation, treating, or swabbing differential pressures will be high
it is desirable to pull the tubing without unseating the packer
it might be desirable to convert the packer to a temporary or permanent
bridge plug
high bottomhole temperatures exist
tubing operating stress variations would not be accommodated with a
retrievable packer without making it impossible to pull
a retrievable packer would have an inadequate bore
Permanent-Retrievable Packers

A recent arrival, this type of packer has the same characteristics as the permanent
packer, but it can, when desired, be released with a special pulling tool and
recovered.
Inflatable Packers
These are packers with a flexible sealing element that can be expanded hydraulically
using either completion fluid or cement. They are used as openhole packers, or when
the casing is buckled or collapsed, preventing the usage of conventional packers.
Inflatable packers cannot stand high pressure differentials and are generally limited
to special applications, such as drillstem testing.
Packer Failures
The major causes of so-called "packer failures" relate to
use of the packer outside its operating range
unsetting of the packer or seal assembly as a result of pressure or
temperature changes
using or setting the packer incorrectly
the packer being in poor condition when run
Difficult operating conditions require more expensive equipment and more rigorous
design work. It is important to remember that a packer is effectively a piston within
the casing and will therefore be heavily affected by changes in differential pressure.
Pressure from below or increasing string tension due to tubing contraction, or both,
tend to unset the following:
weight-set packers
hydraulically set packers not equipped with holddown buttons
locator seal assemblies or overshots
Pressure from above tends to unset tension-set packers or cause collapse-type
failures or leakage at seal assemblies under high pressure differentials.

Tubing/Packer Forces and Movement


Changes in temperature and pressure inside and outside the tubing affect the length
of the tubing string (if the string is designed to permit motion) and the forces
exerted at the packer (if no motion is permitted).
These changes in tubing length or force between producing and pump-in conditions
can be large, and should be considered in choosing a packer. This is especially
important in high temperature (usually deep) wells, and may limit the use of

retrievable packers. Design of slip joints and seal assemblies must also consider
these forces.
(The following section draws heavily on the work of D.J. Hammerlindl (1977) and
Arco as published in JPT February 1977, and on that of Arthur Lubinski, W. S.
Althouse, and T. L. Logan (1962).)
Factors Causing Packer Forces or Tubing Movement
If the tubing string is free to move, its length will change as a result of temperature
and pressure influences, which may be subdivided into thermal, piston force,
ballooning, and buckling effects. Consideration of these effects will determine the
seal length required and/or slip joint design.
If the tubing is anchored to the packer, these effects will result in a change In the
axial tension in the tubing. This can affect not only the design of the uppermost
tubing joint, but also packer shear pin rating and the degree of buckling above the
packer and therefore the through-bore access. A mechanical force is also involved in
this situation.
To consider these effects, it is necessary to define the critical conditions to be
examined. These normally include:
landing conditions;
operating conditions;
shut-in conditions;
killing/stimulating conditions;
pressure test conditions.
Initial calculations are made for a set of assumed landing conditions, e.g.,
-50,000 to +50,000 lb (-225 to +225 kN) tension (-) or compression (+). For convenience a
designer may sometimes choose to examine only the differences between what are considered to
be the most severe conditions (i.e., hot producing to cold stimulating); however, these are not
always apparent (e.g., pressure test loads can be the most severe).
Mechanical Forces
Mechanical forces can be subdivided into tension and compression. Tension results in
the stretching of tubing string. The elongation due to tension forces can be
determined by Hooke's law, which states that the change in length is directly
proportional to the applied force. The equation for Hooke's law in this application is
as follows:

(16)
where:

= length change (inches)


L = tubing length (inches)
Ft = tensional force (-) (lbf)
E = Young's modulus (30 106 psi)
As = cross-sectional area of the tubing wall (in2)
This relationship is the basis of the tubing stretch tables and graphs published by various
equipment manufacturers.
Temperature or Thermal Effects
The length change due to change in temperature is equal to the length of the tubing,
times the coefficient of thermal expansion for steel, times the change in average
temperature:

Since
(17)
then
(18)
where:
= change in tensional force in tubing at the surface due to temperature
change
As the coefficient of thermal expansion for steel () is 6.9
l0-6/F (12.42
l0-6/C) and Young's
modulus (E) is 30 x 106 psi (207 GPa):
= -207 As T lb/F
Ft = -2.57 As T N/C
In most cases, it is adequate to deal with the change in the average string temperature. It is often
assumed that
the completion is at geothermal gradient when landed (1 to 2 F/100 ft, or 1.8 to 3.6
C/100 m)
during stimulation/killing it will stabilize within 20 F (10 C) of ambient
temperature
The temperature during producing conditions is a function of flow rate, gas expansion,
geothermal gradient, thermal insulation of the tubing, and the like. Various computer programs
are available to calculate this in critical cases, but it is usually adequate to use data from offsets
producing under similar conditions. Production test data should be used judiciously since test
rates and flow periods are often too low and too short for thermal stabilization to occur. As a first
approximation designers will often assume a flowing gradient for high rate wells of 0.4 F/100 ft,
or 0.7 C/100 m.
Piston Force Effects
The most familiar form of piston force is that of the stretch and/or stress induced
when making an internal pressure test against a plug set inside a string of tubing.

The force against the plug is equal to the pressure applied times the cross-sectional
area of the tubing ID:
Fp = pt Ai (19)
For example, a 10,000 psi (69,000 kPa) pressure test on 3 1/2-in. (89 mm) tubing (I.D. 2.992 in.)
should result in a force of
Fp = 10,000
(2.992)2
= 70,000 lb (311,000 N)
Where the tubing is inserted into a packer we must similarly consider the piston effects on the
cross section of the steel as it is affected by changes in the internal and external pressure.
The piston force at the packer related to a change in the inside and outside pressures
is
Fp = (Ap - Ai)

pi - (Ap - Ao)

po

(20)

where:
Ap = area of packer bore
Ai = area of tubing ID
Ao = area of tubing OD
The change in tubing length related to this piston force is
(21)
where the pi and po changes are considered positive if they correspond to an increase and
negative if they correspond to a decrease in pressure ( Figure 1 , Schematic showing piston force
at the packer related to a change in the inside and outside pressures).

Figure 1

The piston forces are, in essence, the change in the buoyant force on the tubing.
Ballooning Effects
As the pressure inside the tubing increases, the pipe expands radially. This will cause
an axial shortening. Application of pressure to the annulus will cause the tubing
diameter to contract (reverse ballooning). This will result in a tubing elongation.
The length change accounts for the change in radial pressure forces due to surface
pressure changes (pis and pos) and fluid density changes (i and o) as well as
flow inside the tubing ( ). The formula for calculating the change in length due to
ballooning is

(22)
where:
= Poisson's ratio of the material (for steel = 0.3)
R = ratio OD/ID of the tubing
= drop in pressure in the tubing per unit length due to flow. The
pressure drop is positive when the flow is downward and zero when
there is no flow.
i = density of fluid inside tubing
o = density of fluid outside tubing
pis = surface tubing pressure
pos = surface annulus pressure
Buckling Effects
Buckling is caused by two effects: applied longitudinal compression loads and
internal pressure.
The first is easy to understand as a logical consequence of the loading of a column.
However, the buckling of a pipe under tension as a result of internal pressure is a
difficult concept to appreciate. One way to visualize this type of buckling is to
consider a slightly banana-shaped tubing joint subject to internal pressure ( Figure 2
and Figure 3 , Causes of buckling).

Figure 2

There will be a small unopposed area upon which the pressure will act to cause
distortion of the pipe.

Figure 3

The resulting buckling will reflect the balance of strain and bending energy. This is
similar to having a "fictitious force" (Ff) acting on the end of the pipe.
To better understand buckling resulting from compression loading, consider a string
of tubing freely suspended inside the casing. Now consider an upward force, F,
applied to the lower end of the tubing. This force compresses the string and buckles
the lower portion of the string into a helix. The neutral point is where the buckling
stops. This force decreases with increasing distance from the bottom of the string
and becomes zero at the neutral point. The distance, n, from the bottom of the
tubing to the neutral point is calculated from the following formula:

(23)
where:
W = Ws + iAi - oAo, representing the buoyed weight of the tubing per unit length
F = the force applied to the lower end
It should be understood that in the presence of fluids the neutral point is not the point at which
there is neither tension nor compression, but it is the point below which the string is buckled and
above which the string is straight.

If the neutral point is within the string, then the shortening of the string due to
buckling (Lb) is as follows:

(24)
where I is the moment of inertia of tubing cross section with respect to its diameter:

I=

(D4 - d4)

When the calculated value of the neutral point is above the upper end of the string,
the entire string buckles into a helix.
When the pressure inside the tubing (Pi) is greater than the pressure outside (po) at
the packer, a shortening will occur due to helical buckling. It was determined by
Lubinski, Althouse, and Logan (1962) that, as far as the buckling is concerned, the
tubing behaves as if subject to the following "fictitious" force:

Ff = A (p - p ) (25)
Because part of this force, Ff, appears to be nonexistent, the entire force, F, was given the name
p

fictitious. Further, it was determined that the string would buckle if Ff is positive and would remain
straight if Ff is zero or negative. Substitution into Equation 24 gives

(26)
where Lb is the change in length with respect to the length of the tubing when landed with p i

If the pressure outside the tubing is greater than the pressure inside the tubing at
the packer (po > pi), there is no helical buckling due to pressure.
The total change in tubing length as a result of these various factors (mechanical and
piston forces, and thermal, ballooning, and buckling effects) may be expressed as

L = Lm + Lt + Lp + LB + Lb (27)
The effect of this net change in tubing length on the tubing stress will depend upon the amount of
motion permitted by the packer design.
Permanent Corkscrewing
If the buckling results in the yield strength of the tubing being exceeded, permanent
corkscrewing can occur. In addition to the stresses of helical buckling, the tubing is
subject to elongational stresses, as well as tangential and radial stresses due to
pressure inside and outside the tubing. A triaxial stress analysis must be made where
conditions suggest that significant tubing stress will result. All the major equipment
suppliers have computer and hand calculator programs available for making these
computations. This service is usually available free to purchasers. The production
engineer should spend some time determining the pressure and temperature
conditions to be expected. Typically, a series of calculations is made for various

landing assumptions and the results plotted for analysis. It is also useful to make at
least one hand calculation as a cross-check on the data received.

Exercise
1.. Oilfield Units
A 10,000-ft, high-rate oil well is completed with 5 1/2-in., 15.5-lb/ft tubing (wall
thickness 0.275 in.). Under producing conditions the flowing temperature gradient is
0.4 F/l00 ft and under static conditions the geothermal gradient is 1.8 F/100 ft
from a mean surface temperature of 40 F. When the well is killed with a large
volume of 40 F seawater, the bottomhole temperature drops to 70 F. If free to
move, what tubing movement can be expected from the landing condition to the hot
producing and to the cold injection conditions? If a hydraulic packer were to be used
and set in 30,000 lb of tension, what would be the tension loading on the packer
after killing the well? (Use Equation 18 and ignore piston, ballooning, and buckling
effects.)
(18)

SI Units
A 3050-m, high-rate oil well is completed with 140-mm, 23.1-kgm/m tubing (wall
thickness = 7 mm). Under producing conditions, the flowing temperature gradient is
0.73 C/100 m and under static conditions the geothermal gradient is 3.28 C/l00 m
for a mean surface temperature of 5.0 C. When the well is killed with a large
volume of 5.0 C seawater, the bottomhole temperature drops to 21.0 C. If free to
move, what tubing movement can be expected from the loading condition to the hot
producing and to the cold injection conditions? If a hydraulic packer were to be used
and set in 133.45 kN tension, what would the tension loading on the packer be after
killing the well?
(Use Equation 18 and ignore piston, ballooning, and buckling effects.)

Solution
Oilfield Units
Reservoir temperature = 40 + (1.8 x 10,000 100) = 220 F
Temperature loss up the tubing = 0.4 x 10,000 100 = 40 F
Flowing tubing head temperature = 220 - 40 = 180oF
From landing conditions to producing conditions

Lt = (L) ( ) (

T)

= (10,000) (6.9
where

10-6) (

T)

T=

Landing

= 130F

Producing

= 200F
T=

= 200 - 130 = 70F

Lt = (10,000) (6.9 x 10-6) (70) = +4.83 ft


where:
1

= initial average temperature in tubing string

2 = final average temperature in tubing string


Tsurf = surface tubing string temperature
TBH = bottomhole tubing string temperature
From producing to injection conditions:

Producing

Injection

= 200 F

= 55 F

T=
= 55 - 200 = -145F
Lt = (L) ( ) ( T)
= (10,000) (6.9
10-6) (-145)
= -10.00 ft
Therefore, the maximum overall length change is -10 ft, from producing to injection.
Equation 18 states the change in tension to be
Ft = -E( ) (As) ( T)
for tubing with OD = 5.5 in., and ID = 4.95 in.:

= 4.51 in2
Between landing and injecting conditions, the apparent force acting on the tubing to cause the
contraction effect is

Ft = (-E) ( ) (As) ( T)
= -(30 x 106) (6.9
10-6) (4.51) (55 - 130)
= +70,017 lb force
To maintain the original tubing length, the packer must exert an equal and opposite force:
Fp = -70,017 lb f (tension)
If the tubing was landed in 30,000 lb tension, the net tension at the packer level is
Fp = -70,017 - 30,000
Fp = -100,017 lb force (tension)
This is well in excess of feasible shear pin arrangements and therefore a hydraulic packer cannot
be used. This well should be completed with a permanent packer and a locator seal assembly of
seal receptacle permitting 10 ft of travel.

SI Units
Reservoir temperature = 5 +

= 105.0 C

Temperature loss up the tubing (flowing) =

0.73
= 22.3 C
Flowing tubing head temperature = 105.0 - 22.3 = 82.7C
From loading conditions to producing conditions
Lt = (L) ( ) (

T)

= (3050) (12.42
where

T=

l0-6) (

T)

Landing

= 55C

Producing

= 93.9C
T=

= 93.9 - 55 = 39.9C

Lt = (3050) (12.42 x 10-6) (38.9) = 1.47 m


where:
1

= initial average temperature in tubing string

2 = final average temperature in tubing string


Tsurf = surface tubing string temperature
TBH = bottomhole tubing string temperature

From producing to cold injection conditions

Producing

Injection

= 93.9C

= 13.0C

T = 2 - 1 = 13.0 - 93.9 = -80.9C


Lt = (L) ( ) ( T)
= (3050) (12.42
10-6) (-80.9)
= -3.06 m
Therefore, the maximum overall length change is 3.06 m from production to injection.
Equation 18 states the change in tension to be
Ft = -E( ) (As) ( T)
for tubing, with OD = 140 mm, and ID = 126 mm:
As =
(1402 - 1262) mm2 = 2.925
10-3 m2
Between landing and injection conditions the apparent force acting on the tubing to cause the
contraction effects is
Ft = (-E) ( ) (As) ( T)
= -(206.8
106) (12.42
10-6) (2.925
10-3) (13.0 - 55.0)
Ft = +315.53 kN
To maintain the original tubing length, the packer must exert an equal and opposite force:
Fp = -315.53 kN
If the tubing was landed in 133.45 kN tension, the net load at the packer is
Fp = -315.53 - 133.45
= -448.98 kN tensile force
This is well in excess of feasible shear pin arrangements and therefore a hydraulic set packer
cannot be used.
This well should be completed with a permanent packer and a locator seal assembly
or seal receptacle permitting 3 m of travel.

Artificial Lift Equipment


Anchors
Unanchored tubing in a rod-pumped well will be subject to constant movement. The
tubing will buckle on the upstroke and stretch on the downstroke. This is sometimes
called "breathing." This movement leads to wear and fatigue problems and can result
in inefficient use of the available pumping unit stroke. Tubing anchors are used to
minimize this movement.
Tubing anchors are classified as follows:

tension anchors, which permit the tubing to elongate but not to shorten
compression anchors, which permit shortening but not elongation
fixed anchors
Compression anchors reduce the breathing problems but do not prevent buckling and
are therefore rarely used. Tension anchors, which gradually "walk down" the inside of
the casing as the pump starts to operate and then set the tubing at its maximum
elongation, may damage the casing by repeated slight movement. Therefore, most
anchors used today are of the fixed type (often miscalled tension anchors) ( Figure 1 ,

Figure 1

Mechanically set anchor, Figure 2 ,

Figure 2

Dual hydraulically set anchor, and Figure 3 , Single hydraulically set anchor).

Figure 3

There are two main setting mechanisms:


rotation set
hydraulic set
It is best to select an anchor with a back-up retrieving mechanism so that if the
primary one (i.e., rotation) fails, a secondary release (i.e., shear pins) can be used.
Some hydraulic anchors depend only upon the differential pressure between the
tubing and casing. These have an additional application in preventing seal assemblies
or packers from becoming unlatched during stimulation operations.

Bottomhole Pumps
Details of the bottomhole pump for rod-pumped wells are set out in API Spec 11AX,
which includes a 12-character code to specify each pump type. The most critical is
the second group which is the pump bore (125 to 275 corresponding to 1 1/4 in. to 2
3/4 in. or 32 to 70 mm) and the pump type (R-rod and T-tubing).

The pump displacement in oil field units can be obtained from


PD = Ep

0.1166 S Es N D2

(28)

where:
S = stroke (in.); say, 74 in.
Ep = pump efficiency; assume 90%
N = pump rate; say, 20 SPM
Es = stroke efficiency (or rod stretch); assume 80%
D = pump bore; say, 2 in.
SN < 1500 in./minute (maximum desirable for rod fall)
Using the numbers given above, the pump rate would be
PD = 497 B/d (79 m3/d)

Downhole Completion Accessories


Seating Nipples
There are three main types of seating nipple used as integral parts of the tubing
string:
pump-seating nipples
selective landing nipples
nonselective or no-go landing nipples
Seating nipples, which are used to accommodate a pump, plug, hanger, or flow
control device, consist of a polished bore with an internal diameter just less than the
tubing drift diameter. Usually a lock profile is also required, especially for landing
nipples. Heavy duty tubing sections, called flow couplings, are often run on either
end of a seating nipple to minimize the effects of turbulence ( Figure 1 , Landing
nipple and flow coupling installation).

Figure 1

Seating nipples and the devices that are set inside them are used for the following
purposes:
to facilitate pressure testing of the bottomhole assembly and tubing
couplings, and the setting of hydraulic packers
to land and seal off a bottomhole pump (pump seating nipple)
to isolate the tubing if it is to be run dry for high draw-down perforating
to land wireline retrievable flow controls, such as plugs, tubing safety valves,
bottomhole chokes, and regulators
to plug the well if the tree must be removed
to land bottomhole pressure bombs
to pack-off across blast joints
to install a standing valve for intermittent gas lift

to plug the tailpipe below packer in order to pull the tubing without killing
the well
to temporarily plug the well while the rig is moved on or off the well
Selective Landing Nipples
Selective landing nipples are nipples with a common internal diameter. In some, the
lock profile is varied for easy identification ( Figure 2 ,

Figure 2

Figure 3 ,

Figure 3

Figure 4 and Figure 5 , Landing nipples and locking mandrels).

Figure 4

Others are accessed by tripping the lock mechanism at the selected depth.

Figure 5

Selective nipples are used when more than one nipple is required within a single
string of tubing, and the designer wishes to maintain maximum throughbore. They
should be no closer than 30 ft (10 m) from a similar profile, and at least 10 ft (3 m)
from any change in diameter.
No-Go Landing Nipples
No-go landing nipples are designed with an ID that is slightly restricted to provide a
positive shoulder to locate a locking mandrel. The ID of these nipples should be
checked against the dimensions of any through-tubing equipment that may be used.
This type of nipple is usually located at the bottom of the tubing string or tailpipe and
at least 5 ft below any profile change.
In tailpipe installations, it is best to include a sliding sleeve above the nipple in case
debris prevents the pulling of any plug set in the nipple by regular wireline methods.
Alternatively, a mechanical perforator may be used to punch a hole above the plug.
Sliding Sleeves

Also referred to as sliding side doors or circulating sleeves, these tubing components
are used to obtain access from the tubing to the tubing/casing annulus either for
fluid circulation or to permit a previously isolated zone to be produced ( Figure 1 ,
Sliding sleeves). They are opened and closed with a wireline tool that has a locating
key that engages the profile in the sleeve. A TFL version is also available for subsea
completions.

Figure 1

These devices are typically placed above each packer in the well. Obviously they are
an essential requirement of multizone completions scheduled for selective
production. Many producers run sliding sleeves in each string in a multistring
completion to increase production flexibility.
A sleeve above the upper packer is particularly useful for the following operations:
kick-off by displacing the tubing contents with a low density fluid, thereby
avoiding the use of coiled tubing within the tubing
well killing prior to a tubing pulling job or workover

circulating out completion fluid with a packer fluid (e.g., from mud to brine
or from water to inhibited brine)
testing of subsurface safety valve (SSSV)
temporarily producing a selective zone into the tubing so it can be tested or
so a bottomhole pressure survey can be obtained.
The quality of the elastomer seals in sliding sleeves has improved greatly over the
last decade. They are now much easier to open and less prone to failure. Special
elastomers are needed for some well fluids and suitable design procedures are now
available for elastomers.
A ported nipple is sometimes used in place of a sliding sleeve, although this makes it
necessary to pull the tubing string in order to stop annular access. Alternatively,
some completion engineers prefer to use a side pocket mandrel and valve as a
circulation point above the packer. Note, however, that side pocket mandrels offer a
reduced area to flow and restrict circulation rates.

Side Pocket Mandrels


Side pocket mandrels are a special eccentric nipple that can accommodate a valve in
parallel to the tubing to control access to the annulus ( Figure 1 , Side pocket
mandrels). They are used to install wireline retrievable gas-lift valves, circulation
devices, flow control valves, and injection valves.

Figure 1

The location of side pocket mandrels for gas-lift valves will be determined by the lift
gas pressure available and kickoff requirements.
It is highly desirable to have one or two mandrels located just above the top packer
in high pressure gas well completions. These are used to facilitate a controlled
circulation kill in the event the sliding sleeve is inaccessible or if corrosion-inhibitor
injection is required. Inhibitor may be supplied either through the annulus or via a
special control line, continuously or in batch treatments. Some operators also use
side pocket mandrels to install a pressure and temperature sensor that can transmit
data to the surface via a cable attached to the outside of the tubing.
Some engineers prefer to use a side pocket mandrel instead of a sliding sleeve above
the top packer, since the elastomer seals on a side pocket circulation valve are easily
retrieved and redressed using wireline, while repair of those in a sliding sleeve
requires a workover. However, most circulation valves have a limited throughput
capacity (0.5 b/m or 5 m3/hr) and some operators therefore have a tendency to pull
the valve to increase circulation capability. This can result in a cutting out of the
valve seat in the mandrel, which inevitably requires a workover to replace the
mandrel.

Blast Joints and Flow Couplings


Blast joints and flow couplings are special joints having the same nominal ID as the
tubing, but a greater OD. They are usually manufactured from special heat-treated
steel ( Figure 1 ,

Figure 1

Blast joints, Figure 2 ,

Figure 2

Polished nipples, and Figure 3 , Schematic of polished nipple run to provide sealing
surface in case of blast joint erosion).

Figure 3

While they do not prevent erosion from occurring, their greater thickness can delay
the time to erosion-caused failures.
Blast Joints
Blast joints are used to increase the abrasion resistance of the tubing string against
the jetting action of a producing formation. Blast joints should be located in the
tubing string opposite all upper perforations spanned by the tubing. Blast joints
should also be used in the wellhead area where abrasive fracturing fluids may be
pumped into the casing access. Polished nipples are sometimes included in the
tubing string on either end of a blast joint in order to provide sealing surfaces for a
spacer pipe should the blast joint fail.
Flow couplings
Flow couplings should be run immediately above each selective or no-go landing
nipple in the tubing string that may be used to locate a flow control device. In high
rate or corrosive gas wells, flow couplings should be used above and below all upsets
or profile changes to reduce erosion, especially if the turbulent fluid contains abrasive

particles. Since most flow controls restrict the tubing ID, the tubing above and below
the controls should be protected by use of a flow coupling.

Subsurface Safety Valves (SSSVs)


Application of Downhole Safety Valves
An SSSV must be installed in all offshore wells capable of flow and at onshore
locations in high pressure or sour gas wells in close proximity to housing, public
roads, or rock slide areas. These requirements are often dictated by government
regulations and/or corporate policy.
The objective is to provide a downhole shut-off that will limit the magnitude and
consequences of the hydrocarbon emission if the primary well control device at the
surface (e.g., Christmas tree) is damaged or cannot be operated. This could occur if
a platform was damaged by a storm or major vessel impact, an explosion, a blowout,
or by foundation instability. Similarly, on land a landslide or vehicle impact might
knock off the well-head. There are several different types of subsurface safety
valves.
Flow-Controlled Safety Valves
These are usually deep-set valves whose operation is directly controlled by the well
stream. They are normally wireline retrievable since they must be reset from time to
time, especially as well conditions change. They are designed to be open normally,
but to snap shut if the tubing pressure dips below a threshold or the production rate
exceeds a preset limit ( Figure 1 and Figure 2 , Flow-controlled safety valves).

Figure 1

Figure 2

They normally use spring tension to hold the valve open. The flow passes through a
flow tube containing a bean. If the pressure drop across the bean exceeds the spring
tension the valve will snap closed. The valve can be a ball, flapper, or stem type. The
safety valve is reopened by raising the pressure on the downstream side in excess of
the closed-in bottomhole pressure. Obviously, setting this type of valve requires an
accurate knowledge of well behavior, temperature, and flow conditions.
The major advantage of these valves is that they are cheap and can be set deep in
the well below the packer, protecting both the tubing and annulus. The main
disadvantages are the servicing and design requirements, the restrictions to flow
capacity and flexibility, and the risk of inadvertent reopening as a result of fluids lost
into the wellbore (e.g., seawater or mud in the event of an offshore collision of a
boat with a wellhead).
Surface-Controlled Subsurface Safety Valves (SCSSSVs)
The SCSSSV is a fail-close valve that is held open by a high pressure control line (
Figure 1 , Tubing-retrievable, surface controlled, subsurface safety valve).

Figure 1

If the control line is severed in an event that damages the tubing or wellhead,
pressure will leak off and the safety valve will close. The control line is generally
connected to an emergency shutdown system to give automatic closure during
unsafe or alarm conditions (such as fire or detection of gas). There is usually a
control panel with pressure gauges and control valves for all of the wells on an
offshore platform. Surface-controlled valves are the type of downhole safety valve
most commonly favored today. In some countries, the regulations require the use of
this type of valve in all offshore wells and onshore sour wells capable of flow. There
are two basic types of SCSSSV: wireline retrievable (TFL retrievable) and tubing
retrievable.
With the wireline retrievable valve, an SSSV landing nipple is installed in the tubing
string. This is basically a landing nipple with a port through which the control line
enters between a set of packings on the SSSV. The SSSV is installed across this
nipple. This type of valve can have a service life of 18 to 24 months or longer,
although many valves fail during periodic testing. The relative ease of servicing and
replacing wireline retrievable valves therefore offers distinct advantages. The main
disadvantages of wireline retrievable SSSVs are that
the service life is short

the restricted throughbore means the valve has to be pulled for deep
wireline or through-tubing work
the turbulence in the flow stream increases pressure loss and erosion
problems
we are forced to rely on a lock to make sure the valve is not blown out of
the well on closure
On the other hand, the initial costs are relatively low and servicing can be
undertaken with minimal disruption to production.
The tubing-retrievable SSSV valve is an integral part of the tubing string. As a result,
it generally has a larger through-bore than the retrievable-type valves and may even
be designed with an internal diameter that is the same as the tubing (a fullbore
valve). Also, it is not so dependent upon elastomer seals and therefore has a much
longer service life (5 to 20 years, depending on design and materials selected) .
Tubing-retrievable valves with all metal-to-metal seals are available for severe
environments.
Since most valve failures are caused by elastomer problems and since tubingretrievable valves require a rig entry for repair work, these valves are often backed
up with a nipple section positioned to accept a wireline-retrievable valve. Use of this
valve insert minimizes the impact of a valve failure on production. This feature is
incorporated in tubing retrievable valves together with a lock-out sleeve for the
original valve, to lock it permanently open. Service procedures have been developed
for installing the insert valve using both TFL and wireline techniques.
To reduce the number of critical seals, many companies prefer single-control line
valves. Although spring design does limit the depth to which this type of valve can be
set, technology improvements have pushed the limit from around 650 ft (200 m) to
in excess of 3250 ft (1000 m). Balance-line valves (valves with two control lines, one
to close and one to open), which were developed to overcome the earlier depth
limitations, are therefore becoming less popular. However, they have an advantage in
that they can be pumped closed to facilitate the cutting of an obstruction in the
valve, such as wireline.
A balance-line valve can also be set at any depth, the critical issue being the closing
time and control line efficiency. Protection systems for the control lines have been
considerably improved over the last decade so that installation damage is much less
common. New hydraulic fluids have also reduced control line pressure losses, so that
it is possible to set a balance-line valve at the packer level. Field trials are also in
progress on the use of electrical control systems for deep-set surface-controlled
subsurface safety valves.
The SCSSSV may incorporate a ball valve or a flapper valve. Ball valves are often
considered more robust and can sometimes cut wireline when they are closed with it
across the valve. However, they are prone to damage by sand and improper
operation. The simpler flapper valve has the advantage of always being reopenable
(mechanically if necessary) should it become stuck in the closed position. Extensive
studies have shown that flapper valves are more reliable than ball valves; as a result,
most operators run flapper style SCSSSVs. Operators also have the choice of running

tubing-retrievable SCSSSVs with lockout option. These valves offer redundancy (i.e.,
the ability to lock open one valve and run a second valve inside). Sleeve-type valves
are also available.
Setting Depth
The recommended setting depth for a safety valve is often a matter of company
philosophy and operating procedure. Where concurrent drilling and production is
undertaken, many companies like to set the safety valve beyond the kickoff point so
that it can be used to shut in the well during the tophole drilling and kickoff of an
adjacent well. Other designers prefer not to subject the valve to significant bending
and therefore favor its installation nearer to the surface. Subsurface safety valves
are normally set at least 150 ft (50 m) below the surface or sea floor.
Bottomhole Chokes and Regulators
These are not safety devices in the strict sense of the word; however they are often
used with the objective of limiting the rate at which a well could produce and
therefore blowout, limiting the surface operating pressures in high pressure wells,
and limiting the drawdown rates on wells that have a tendency to produce sand.
They are also used to take maximum advantage of the formation temperature to
avoid hydrate formation during choking of a high pressure gas stream at surface
temperatures. Chokes are designed to give a constant rate, while regulators give
constant choking or pressure differential.
The other safety device that is often used in injection wells is a simple check valve
that will permit injection but not production. This device may be used to prevent
backflow of water and formation sand, should injection cease.
All of these devices are normally wireline retrievable.

Wellhead Equipment
Wellheads
Wellheads typically are the joint responsibility of the production department (tubing
head and Christmas tree) and drilling department (casing head and intermediate
casing head). Figure 1 shows a typical flanged wellhead.

Figure 1

The size and pressure ratings of wellheads are dictated by the design considerations
for the tubulars (e.g., tubing size, casing size, kill and stimulation pressure
requirements, flowing pressure requirements). However, government regulations
sometimes require that the rating of the upper part of the wellhead be at least equal
to the reservoir pressure.
Wellhead specifications are presented in API Spec. 6A. The standard wellhead ratings
are 1000, 2000, 3000, 5000, 10,000, 15,000, and 20,000 psi (7, 14, 21, 34, 69,
103, 130 MPa).
Wellhead components are generally flanged, although threaded components are
permitted on low pressure wellheads with pressures less than 2000 psi (14 MPa).
Threaded valve and choke connections can be used with pressures of up to 5000 psi
(34 MPa) in wells less than 12,000 ft (3700 m) deep, but are not recommended.
Clamped connections are sometimes used in the intermediate pressure range 2000
to 10,000 psi (14 to 69 MPa) ( Figure 2 , Wellhead and Christmas tree for a dualtubing completion utilizing clamp-type connections). For wells producing H2S gas, the
wellhead materials must conform to NACE specifications.

Figure 2

Tubing Heads and Hangers


The tubing head packs off around the production casing ( Figure 1 , Tubing head and
tubing hanger installation).

Figure 1

It should have an outlet for access to the tubing/casing annulus, with an internal
thread to receive a plug when redressing of the side outlet valve is necessary. The
rating of the upper flange must be compatible with the Christmas tree. The tubing
head should have lock-down screws for the hanger, and the lower flange size and
rating must be compatible with the casing head flange.
The bore and size of the top flange are generally determined by completion and well
servicing requirements (BOP size, packer, and tool ODs) rather than the Christmas
tree flange size.
Like the casing, the pressure rating of the tubing head spool is often dictated by
stimulation pressure requirements and may therefore be of a higher rating than the
Christmas tree, which can be removed or protected during stimulation.
Offshore, a compact wellhead, or unihead, is often used to combine both the casing
and tubing spool's function and reduce the overall height of the wellhead.
Three types of tubing hangers are commonly used:

the boll weevil (also called a threaded mandrel) hanger, which is an integral
part of the tubing string and therefore a fixed point that shoulders into the
tubing head spool ( Figure 2 , Boll weevil tubing hanger)

Figure 2

the wrap around hanger, which is hinged to permit installation onto any part
of the tubing other than a coupling
the dual hanger, either multibore mandrel or split hanger
The mandrel types are the most common.
It is highly desirable to have an internal thread in the tubing hanger to allow the
installation of a back pressure valve while removing, repairing, or pressure testing
the tree. This can be installed and removed under pressure with a special tool.

Christmas Trees
(see Figure 1 , Typical flanged wellhead)

Figure 1

There are three main types of trees: the assembled tree, the solid block tree, and
the control head tree (often found on thermal wells). The major components (from
bottom up) are
the flange
the master valve(s)
the tee or flow cross
the swab valve
the crown plug
the wing valve
the bean box or choke

the flow line valve


For high rate wells the flow tee is often Y-shaped to reduce turbulence and erosion. Similarly, a
flow control valve may be installed in a straight run rather than in the conventional right-angled
bean arrangements.
A second side outlet is often used on high pressure wells as a connection for a tubing
kill line. Similarly, two master valves are often used in severe operating conditions.
This is often a regulatory requirement in sour or high pressure wells.

Tubing
Size

Rating

Tree
Bore

Tree
Drift

(in)

(mm)

(psi)

(MPa)

(in)

(mm)

(in)

(mm)

2 3/8

(60)

15,000

(103)

2 1/16

(52)

1.901

(48)

2 7/8

(73)

15,000

(103)

2 9/16

(65)

2.347

(60)

3 1/2

(89)

5,000

(34.4)

3 1/8

(79)

2.867

(73)

3 1/2

(89)

15,000

(103)

3 1/16

(78)

2.867

(73)

Table 1: Christmas tree specifications


Full opening gate valves are used for the master and swab valves
( Figure 2 ,

Figure 2

Manually-operated gate valve, Figure 3 ,

Figure 3

Pressure-actuated gate valve in open position, and Figure 4 , Pressure-actuated gate


valve in closed position).

Figure 4

These should not be opened when a significant differential pressure exists across the
closed valve.
The throughbore of the tree is specified by the API and is generally 1/16 in. larger
than the tubing ID to facilitate installation of a back pressure valve in the tubing
hanger. Tree sizes are shown in Table 1.
Although the body of a Christmas tree is normally pressure tested to twice the
working pressure for trees rated at 5000 psi (34.4 MPa) and less, and 1.5 times the
working pressure for 7500 to 20,000 psi (52 to 140 MPa) ratings, the flange bolts
and valves may not necessarily have the same rating. Therefore, it is extremely
imprudent to overload Christmas trees when stimulating a well. Similarly, many
valves are unidirectional and this should be taken into account when planning
pressure test sequences. Valve gates can be damaged by applying significant
pressure from the wrong side.
Beans and Chokes
In flowing wells, rate is controlled by a bean, choke, or flow control valve.
Traditionally, the most common was the fixed bean operating under critical flow

conditions (i.e., at sonic velocity). Under these conditions the upstream pressure, or
tubing head pressure (THP) , is independent of the downstream pressure, or flow line
pressure (FLP). To achieve this, THP must be greater than or equal to 2.0 times the
flow line pressure. The advantages of operating under these conditions include the
following:
over the short term (generally one to three months) the well rate is fixed,
and a single monthly test is representative of the entire producing period
test separator conditions need not be the same as the bulk separator to
ensure a representative test, since fluctuations in downstream pressure do
not affect THP at sonic velocity
well flow rate is limited in event of a line break
lower pressure ratings can be used for flow lines and separators
the sand face is not subjected to production surges in event of a production
facility fluctuation (this point is particularly important in weak formations)
choke performance can be used as an indication of production rate
The disadvantages relate primarily to lower pressure wells and gas wells:
the choke introduces a major pressure loss into the system
flow lines may need to be larger to accommodate the higher flow velocities
without excessive erosion or pressure loss
associated cooling can cause hydrate formation at the choke
choke beans are inconvenient for changing production rates in accordance
with changes in gas sales requirements
To meet the last objection, motorized or manual variable chokes or flow control
valves are often used on key wells so that the operator can quickly change the field
flow rate.

You might also like