Sie sind auf Seite 1von 37

Mobil Coiled Module II Mechanical Performance

Tubing Manual

Module II
Mechanical Performance
Common mechanical operations for CT include:
• Logging and drifting

• Perforating

• Setting/removing plugs and packers

• Sliding sleeves

• Drilling

• Milling/scale removal

• Fishing

Forces acting on a CT string govern its performance during such


operations. Determining or predicting that performance requires
knowledge of the well path, tubing and BHA dimensions and properties,
wellbore geometry, densities of the fluids inside and outside the tubing,
pressures throughout the system, and the coefficient of friction (Cf)
between the CT string and wellbore. The following sections explore the
basic concepts surrounding this problem.

General Force
Balance
Figure 9 shows CT and a BHA in a horizontal well with force Fb acting on
the bottom of the BHA. Fb is the resultant of the axial force, i.e., setdown
weight (SDW) or overpull, and torque. The fundamental problem is to find
the resultant force Fs at the top of the CT, because it is the primary real-
time indication of what is happening down-hole.
A secondary indication is a surface display of force measurements at the
BHA. CT “weight” (axial force) measured by the weight indicator at the
surface (CTWsurf) is the sum of the axial component of Fs, stripper friction,
axial pressure force, and reel tension.

July 30, 1996 Proprietary information of Mobil 11


Module II Mechanical Performance Mobil Coiled
Tubing Manual

CT

/
Figure 9:
CT in a horizontal well

BHA

A typical method of finding F, is to divide the CT and BHA into segments


and perform a three-dimensional force balance on each segment from
the bottom up. Segmentation is necessary because boundary conditions
and geometry often change with depth. The objective is to calculate the
force acting at the top of each segment (F,) given the force acting at its
bottom (Fb). Equations 1 and 2 give the axial and torsional components of
the force balance over any segment.

Equation 1: Axial component offorce balance

Axial force at Top = Axial force at Bottom + Weight Component f Axial Drag

Equation 2: Torsional component offorce balance

Torque at Top = Torque at Bottom - Torsional Drag

b@ NOTE: Drag always opposes motion.

Since drag opposes motion, use the (-) sign for running into hole (RIH)
and torque, but use the (+) sign for pulling out of hole (POOH). The forces
acting on each CT and BHA segment could include:

Axial forces due to:


l Buoyed weight
l Frictional drag (in the stripper and against the wellbore)

12 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

l SDW or over-pull
0 Pressure x area
. Tension from the CT reel
l Hydraulic drag due to fluid flow

Normal Forces due to:


l Buoyed weight
0 Curvature
l Buckling

Torsional forces from the BHA due to:


l An orienting device
l Motor (drilling reactive torque)

Segmenting the CT and BHA shown in Figure 9, beginning at the lower


end with segment 1 and ending at the upper end with segment k, produces
the condition shown in Figure 10.

- Segment k
- Segment k-l

F,, -+ lSegment1 +- F, = F,,.,


Figure 10:
CT and BHA divided into segments for a
force balance
Segment 1

t
Segment 2

Given the force acting on the bottom of segment 1 (Fb), the balance of all
of the forces acting on segment 1 yields the force acting at its upper end
(F,t). This force becomes Fb2 , the force acting on the bottom of segment
2. The balance of all the forces acting on segment 2 yields the force acting
at its upper end (FQ). Proceeding upward in this fashion for each segment
in the string gives the force balance on uppermost segment and yields F,.

July 30, 1996 Proprietary information of Mobil 13


Module II Mechanical Performance Mobil Coiled
Tubing Manual

The correct force balance equations on any given segment depend upon its
geometry, curvature, and buckling condition. The following section
describes the force balance procedure for typical segments under various
conditions. The “soft string” model described below neglects effects of
moments on each segment. The sign convention used for axial force is
positive (+) for tension and negative (-) for compression. Torque is always
positive (+). For the sake of simplicity, the following discussion also
ignores effects of pressure from surrounding fluids on force balance. Such
a balance is known as “effective” force balance. Correcting the “effective”
force F,, for pressure effects yields the real force at any point n:
Fir,n(rra/)= Fs,n- %A0 + eA,

Force Balance for


Straight Segments

Consider a straight segment of length AL in the CT string (see Figure 11)


where :

1. The segment is a hollow cylinder

2. The segment rests on a plane inclined at angle 8 from vertical

3. The axial force F 1 and the torque Tt acting at the lower end of the
segment are known.

Figure 11:
Force balance on a straight cylinder
and plane

14 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

The goal is to determine the axial force F2 and torque T2 acting on the
upper end of the segment. W, the buoyed weight per unit length of the
cylinder, depends upon the density of the fluids inside and outside the
cylinder. If a unit length of the cylinder weighs Wair in air, the fluid inside
has density ρi, and the fluid outside has density ρo, use Eq. 3 to calculate
the buoyed weight per unit length.+

Equation 3: Buoyed weight per unit length

Ø W = Wair + ρi Ai − ρo Ao (3)

where
π π
Ao = OD 2 Ai = ( OD − 2t )
2

4 , 4

OD is the cylinder's outer diameter, and t is its wall thickness. W is the


vector sum of two orthogonal components, Wa (axial) and Wn (normal).
The axial component, Wa = W cosθ , enters the axial force balance as
Wa x ∆L. The component normal to the surface of the plane, Wn = W sinθ ,
enters the axial force balance through frictional drag, Cf x Wn x ∆L, where
Cf is the coefficient of friction. Thus, Eq. 4 gives the axial force at the
upper end of the segment.

Equation 4: Axial force at the upper end of the segment

Ø F2 = F1 + W cosθ × ∆L ± Cf × W sin θ × ∆L (4)

If the segment is moving down, as depicted in Figure 11, subtract the drag
term. If the segment is moving up, add the drag term.
Equation 4 is analogous to Eq. 1 where W cosθ × ∆L is the axial
component of weight, and Cf × W sinθ × ∆L is the drag.

+
As used herein, density actually means specific weight and has the units of weight/unit
volume.

July 30, 1996 Proprietary information of Mobil 15


Module II Mechanical Performance Mobil Coiled
Tubing Manual

Similarly, Eq. 5 gives the torque at the upper end of the segment.

Equation 5: Torque at the upper end of the segment

q=lp- xCf xWsinBxAL (5)

Since torque is always positive, torsional drag always reduces the torque at
the upper end of the segment. Due to friction, torque from an orienting
device or motor in the BHA usually dissipates before reaching the surface.
(See Module VI, Minimizing Risk for CT Operations for an example.)

Force Balance for


Segments in Curved
Wellbores

A straight segment forced into a curved wellbore, as shown by Figure 12,


exerts a normal force C, against the wellbore in addition to the normal
component of its buoyed weight. C,, often called the “capstan” force, is
distributed along the line of contact between the segment and the wellbore.
The effective normal force for a curved segment is the vector sum of C,
and the normal force due to weight acting in a direction determined by the
orientation and rate of curvature of the segment.

Ck_ F ‘“VW F2 Figure 12:


Normal force due to curvature of a segment
” 1
Fl

a: Building Curvature b: Dropping Curvature

Thus, a vector force balance, such as in MEPTEC’s CT simulator’, is


required to properly evaluate the effective normal force on a segment in a
curved wellbore. An explanation of this procedure is beyond the scope of
this manual, but Eq. 6 approximates the effective normal force for a
“short” segment in a curved wellbore.

’ CT Simulator Theory Discussion in the CT Simulator Manual

16 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

Equation 6: Approximation of effective normal force for a short segment in a


curved wellbore

Ø
Fn = ∆L × [(F θ ′ + W sinθ )
1
2
+ ( F1φ ′ sin θ )
2
] (6)

where ∆L, W, and F1 are as defined previously, θ is the average


inclination of the segment, θ ′ is the local build rate (i.e. change in
inclination per unit length), and φ ′ (i.e. change in azimuth per unit length)
is the local walk rate. Local build and walk rate have units of radians per
unit length. Eq. 6 gives reasonable results for segments short enough that
the total curvature angle of the segment is less than 1°. Since Eq. 6 reduces
to Fn = ∆L × W sinθ for straight segments ( θ ′ = φ ′ = 0), it can be applied
to any unbuckled segment, curved or straight.
For a given normal force, Eq. 7 and Eq. 8 yield the axial force and torque
at the upper end of a curved segment.

Equation 7: Axial force at the upper end of a curved segment

Ø F2 = F1 + W cosθ × ∆L ± Cf × Fn (7)

Equation 8: Torque at the upper end of a curved segment

 OD 
T2 = T1 −   × Cf × Fn
 2  (8)
Ø
Note the similarity between Eq. 7 and Eq. 1, and between Eq. 8 and Eq. 2.
As before, subtract drag during RIH and add drag during POOH.
Direct measurements of axial force on the BHA from a tension head tool or
auxiliary measurement sond (AMS) can improve the interpretation of
down-hole conditions. Assuming the BHA does not buckle, use these data
and Eq. 7 to calculate Cf as the BHA traverses the wellbore. Such indirect
measurements of Cf are extremely valuable for predicting CT performance
(see Module IV, Spreadsheet Calculations, and Module V, Simulators).

Force Balance for


Buckled Segments
CT is prone to buckling while RIH in deep and/or extended reach wells.
The section “Buckling and Lockup” on page 22 discusses conditions
required to buckle a segment.

July 30, 1996 Proprietary information of Mobil 17


Module II Mechanical Performance Mobil Coiled
Tubing Manual

NOTE: The force balance equations for a buckled segment are different
from those for an unbuckled segment.

The Excel spreadsheet “CT Forces & Stresses”, described on page 55,
contains equations to calculate buckling limits and post-buckling forces.
The following discussion assumes a portion of the CT is helically buckled
and subjected to compressive axial force.

Buckled Segment in a
Straight Inclined Section
of the Wellbore

Figure 13 illustrates a helically buckled segment in a straight wellbore.


Once a segment is buckled, it responds to further increase in axial
compressive force by tending to “expand” outwards. Since it cannot
expand outwards due to the presence of the constraining wellbore, a
normal force, B,, develops. This normal force due to buckling adds to the
effective normal force in the force balance. B, is distributed along the line
of contact between the segment and the wellbore.

Figure 13:
Buckling normal force in a segmenl

18 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

Equation 9 gives the normal force per unit length due to buckling2,

Equation 9: Normal force per unit length due to buckling

rc F 2
Bn = (9)
Ø 4 EI

where F is the net axial compressive force on the segment, EI is the


segment's bending stiffness (flexural rigidity), and rc is the radial clearance
between the hole ( IDhole ) and segment, given by Eq. 10.

Equation 10: Radial clearance between the hole and segment


IDhole − OD
rc = (10)
Ø 2

If the segment is relatively short, replace F with F1 in Eq. 9. Young's


modulus, E, for steel is 30 x 106 psi in customary U. S. units. Equation 11
gives the segment's moment of inertia.

Equation 11: Moment of inertia for a segment

π[( OD 4 − (OD − 2t ) 4 ]
I= (11)
Ø 64

For buckled segments in a straight wellbore, effective normal force is


Wn + Bn , and drag becomes Cf × (Wn + Bn ) . Thus, Eq. 12 and Eq. 13 give
the axial force and torque at the upper end of a buckled segment of length
∆L.

Equation 12: Axial force at the upper end of a short buckled segment in a
straight wellbore

Ø F2 = F1 + W cosθ × ∆L − Cf × (W sin θ + Bn ) × ∆L (12)

2
Mitchell, R.F., 1986, “Simple Frictional Analysis of Helical Buckling of Tubing”, SPE
Drilling Engineering, December 1986, pp. 303-310.

July 30, 1996 Proprietary information of Mobil 19


Module II Mechanical Performance Mobil Coiled
Tubing Manual

Equation 13: Torque at the upper end of a short buckled segment in a straight
wellbore

 OD 
T2 = T1 −   × Cf × (W sinθ + Bn ) × ∆L (13)
 2 
Ø
Equation 12 and Eq. 13 are valid only for short segments. Always subtract
drag for buckled segments, because buckling only occurs while RIH.

! NOTE: Bn is proportional to axial force squared and radial clearance,


while inversely proportional to a segment's bending stiffness.
Therefore, the drag term in Eq. 12 and Eq. 13 increases with the square of
the axial force.

This explains why maximum reach of CT in a wellbore and Set Down


Weight (SDW) or weight on bit (WOB) at a given depth are so sensitive to
buckling. Decreasing radial clearance and increasing the segment's bending
stiffness (larger OD or thicker wall) will decrease drag on a buckled
segment for a given axial force.

Buckled Segment in a
Curved Section of the
Wellbore
If buckling occurs in a CT segment in a curved section of the wellbore,
total normal force is the vector sum Wn + Bn + Cn . However, the buckling
curvature is usually much more severe than the wellbore curvature so that
Cn << Bn. In other words, the normal force due to wellbore curvature is
much less than the normal force due to buckling. Therefore, Eq. 14 and Eq.
15 give the axial force and torque at the upper end of a buckled segment of
length ∆L in a curved wellbore.

Equation 14: Axial force at the upper end of a short buckled segment in a curved
wellbore

F2 = F1 + W cosθ × ∆L − Cf × (W sin θ + Bn ) × ∆L
Ø (14)

20 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

Equation 15: Torque at the upper end of a short buckled segment in a curved
wellbore

 OD 
T2 = T1 −   × Cf × (W sinθ + Bn ) × ∆L (15)
Ø  2 

where θ is the average angle of inclination over the segment. Equation 14


and Eq. 15 are valid only for short segments.

! NOTE: The drag term in Eq. 14 and Eq. 15 is proportional to the square
of the axial force, while the drag term in Eq. 7 and Eq. 8 varies linearly
with the axial force. All else being equal, drag on a buckled segment is
much greater than drag on an unbuckled segment in a curved wellbore.

Friction Coefficients
Cf for sliding friction is independent of direction unless one or both
surfaces have an oriented structure (not likely for steel). Consequently,
MEPTEC advocates using the same Cf for both RIH and POOH based on
the local surface conditions.
MEPTEC friction research, analyses of tension head data, and numerous
studies reported in the public domain support the following choices for Cf.
• Water-wet steel surfaces, Cf > 0.30

• Lubricated water-wet steel surfaces, Cf > 0.20

• Oil-wet steel surfaces, Cf > 0.15

• Steel on rock, Cf > 0.40

Overall Force Balance


Recall that the force balance so far ignores pressure-area effects. A
pressure differential across the stripper imparts a compressive axial force in
the CT string given by Eq. 16. If Pi is the CT internal pressure and Po is the
CT external pressure (annulus) just below the stripper, Eq. 16 gives the
pressure-area force.

July 30, 1996 Proprietary information of Mobil 21


Module II Mechanical Performance Mobil Coiled
Tubing Manual

Equation 16: Compressive axial force from stripper

 πOD 2 
Parea = Po ×   (16)
 4 
Ø
Assuming CTWsurf accurately reflects an axial force balance over the entire
CT string, Eq. 17 and Eq. 18 give the axial force balance equations for RIH
and POOH, respectively.

Equation 17: Axial force balance for RIH

Ø Slackoff = Buoyed Weight - Drag - P a r e a - S t r i p p e r F r i c t i o n - R e e l T e n s i o n (17)

Equation 18: Axial force balance for POOH

Ø Pickup = Buoyed Weight + Drag - P a r e a + S t r i p p e r F r i c t i o n - R e e l T e n s i o n (18)

where:
Slackoff = CTWsurf during RIH
Pickup = CTWsurf during POOH
Buoyed Weight = axial component of total buoyed weight of CT and BHA
Drag = total drag of the CT string against the wellbore
Stripper Friction = drag on the CT caused by the stripper
Reel Tension = axial force acting on the CT due to the force required to
wrap/unwrap it onto/off the reel

Buckling and
Lockup
Buckling of drillpipe during drilling can have extremely serious
consequences, including rapid fatigue failure of the drillpipe. Buckling of
casing or tubing usually causes high stresses that weaken the tubular and
degrade its pressure rating. Consequently, drillers take great pains to avoid
buckling jointed tubulars. Buckling of CT is common for operations in
deep and/or extended reach wells. In fact, CT normally enters the wellbore
in a "buckled" shape with residual curvature from its plastic deformation on
the reel. This residual curvature is not a buckled state, but helps promote
buckling as axial compressive force on the CT increases.

22 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

! NOTE: By itself, buckling of CT is not a serious problem and does not


damage the tubing.

Under favorable conditions, buckled CT can continue to slide and transmit


axial force. However, Eq. 12 and Eq. 14 show that buckling significantly
increases drag between the CT and wellbore. This always reduces
available WOB or SDW, and may lead to lockup, one of the most serious
problems for CT operations. The following section summarizes important
concepts about buckling and lock-up. Appendix A describes MEPTEC
buckling research and contains a detailed discussion of relevant equations.

Buckling Limits
A CT segment inside a wellbore buckles into different shapes when the
effective axial compressive force acting on it exceeds values determined by
the particular combination of geometry and physical properties of the
segment. The segment remains unbuckled for low axial compressive force.
When the axial compressive force increases to the critical sinusoidal
buckling limit, the segment deforms into a sinusoidal or "snake-like" shape
in continuous contact with the wellbore.
The buckled segment does not move way from the wellbore nor lie in a
plane. The segment continues to change shape as the axial force increases
beyond the critical sinusoidal buckling limit, but the normal force exerted
by the segment on the wellbore is due mainly to the weight of the segment.

! NOTE: Sinusoidal buckling does not present a limiting condition for CT


operations but merely signals that onset of instability. In CT mechanics,
helical buckling and post-helical behavior are more important.

If the axial compressive force continues to increase past the critical helical
buckling limit, the segment assumes a helical shape in continuous contact
with the wellbore. For CT segments inclined at angles less than ≅15°, the
pitch of the helix is on the order of 150-200 ft. For highly inclined
segments, the pitch of the helix is on the order of 40-60 ft.
After the segment is buckled helically, the normal force exerted by the
segment on the wellbore gains a component proportional to the square of
the axial compressive force, Eq. 9. Thus, drag on a helically buckled
segment increases rapidly with increasing axial compressive force. Properly
accounting for this additional drag in the force balance, requires knowledge
of when the axial force on a segment exceeds the critical limit for helical
buckling. The Excel spreadsheet "CT Forces & Stresses" (see page 55)

July 30, 1996 Proprietary information of Mobil 23


Module II Mechanical Performance Mobil Coiled
Tubing Manual

calculates helical buckling limits and post-buckling forces for CT segments.


The following section discusses critical buckling limits for a CT segment.

Vertical Segments
Equation 19 gives the critical helical buckling limit for vertical straight
segments and segments inclined at angles less than ≅15°3.

Equation 19: Critical helical buckling limit for near-vertical, straight segments

Ø FVH = 194
. 3 EI × W 2
(19)

where W and EI are as defined earlier. Note that Eq. 19 depends only on
dimensions and properties of the segment. For common CT sizes in vertical
holes, FVH is typically less than 200 lbs compression. This force may seem
insignificant, but for many situations most of the CT in a vertical wellbore
is in tension, and buckling is not an issue. If part of a CT string is in
compression and the remainder is in tension, the location where axial force
changes from tension to compression is called the neutral point.

Inclined Segments
Equation 20 gives the critical sinusoidal buckling limit for an inclined
segment with inclination angle θ4.

Equation 20: Critical sinusoidal buckling limit for an inclined segment

EI × W sinθ
FCS = 2
Ø rc
(20)
Equation 10 defines the radial clearance, rc. The axial compressive force
required to helically buckle an inclined segment is about 41% greater than
FCS . Equation 21 gives the critical helical buckling limit for a straight
inclined segment5.

3
Lubinski, A., Althouse, W.S., and Logan, J.L., 1962, “Helical Buckling of Tubing
Sealed in Packers,” Journal of Petroleum Technology, June 1962, pp. 655-670.
4
Dawson, R. and Paslay, P.R., 1984, “Drillpipe Buckling in Inclined holes,” Journal of
Petroleum Technology, October 1984, pp. 1734-1738.
5
Chen, Y., Lin, Y., and Cheatham, J.B., 1990, “Tubing and Casing Buckling in
Horizontal Wells,” Journal of Petroleum Engineers, February 1990, pp. 140-191.

24 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

Equation 21: Critical helical buckling limit for a straight inclined segment

EI × W sinθ
FCH = 2 2
Ø rc

For common CT sizes, FCH for a horizontal segment can be 20-30 times (21)
the axial compressive force required to helically buckle the same segment
in a vertical position. This partially explains why CT usually buckles first
near the bottom of the vertical portion of a well during RIH. Another
reason is that drag is much higher on curved and inclined segments leading
to higher axial compressive force at the bottom of the vertical section.
Note that FCH increases with decreasing radial clearance and increasing
segment bending stiffness, weight, and inclination. This provides several
options for reducing the tendency of a segment to buckle helically. If
buckling could be a problem, larger diameter CT simultaneously increases I
and W while decreasing rc. Another alternative is to increase the CT wall
thickness, which simultaneously increases I and W. If the CT dimensions
are fixed, the only way to increase the critical helical buckling limit is to
lower mud weight (increase W) or conduct the CT operation inside a
smaller casing or hole size (decrease rc). The Excel spreadsheet "CT
Forces & Stresses" (see page 55) contains equations to calculate buckling
limits and post-buckling forces.

Effects of Curvature on
Buckling

Traditionally, Eq. 21 has been used in curved holes by considering θ as the


average inclination of the curved segment. However, this does not account
for the stabilizing effect of curvature on the buckling. For example, Eq. 21
will give a constant value for FCH in a curved hole regardless of the
direction or rate of curvature. MEPTEC uses a procedure that accounts for
the effect of curvature by including the normal force due to curvature,
Eq. 6, as additional resistance to buckling6. This procedure gives results
consistent with MEPTEC's experimental observations, and with the results
of a rigorous theoretical study of the problem by an independent
consultant7.

6
He, X, and Kyllingstad, A., 1993, “Helical Buckling and Lock-up Conditions for
Coiled Tubing in Curved Wills”, Paper SPE 25370, presented at the SPE Asia Pacific
Oil and Gas Conference, Feb. 1993.
7
Paslay, P.R., “Buckling of a Rod Confined to be in Contact with a Toroidal Surface,
Parts I and II”, MEPTEC Confidential Report, October 1993.

July 30, 1996 Proprietary information of Mobil 25


Module II Mechanical Performance Mobil Coiled
Tubing Manual

Replacing W sinθ , the normal force due to segment weight Eq. 21, with
Fn, the effective normal force from Eq. 6, and simplifying, yields a quartic
polynomial equation for the axial force required for helical buckling, FH :

Equation 22: Axial force required for helical buckling in a curved hole

[( Fθ ′ + W sin θ ) ]
2
 8 EI 
+ ( Fφ ′ sin θ )
2
F = 
4 2

 rc 
H
Ø
Note that Eq. 22 is derived in terms of force per unit length. The required
critical helical buckling limit is one of the solutions of Eq. 22. Since FH is (22)
the axial compressive force (assumed positive here), only positive roots of
Eq. 22 are physically admissible. When only one positive root exists, it is
the required critical helical buckling limit. If multiple positive roots exist,
the selection process depends upon the curvature and the mechanics of
buckling as discussed in Appendix B.

! NOTE: In general, curvature stabilizes a segment against buckling. That is,


CT buckles more easily in straight sections of the wellbore than in doglegs
or build sections.

Comparing Eq. 21 and Eq. 22 yields the following:


• In building curvature, FH > FCH

• For moderately dropping curvature, FH < FCH


• For high dropping curvature, FH > FCH
• For purely azimuthal curvature, FH > FCH

Does the foregoing mean that a tortuous or curvaceous wellbore is


beneficial for CT operations? Reducing the buckled length of CT would
extend its reach into a well. On the other hand, Eq. 6 clearly shows that
curvature causes higher drag which may shift the location of buckling
upwards. Resolution of these conflicting effects generally requires
modeling a CT operation with a good CT simulator (see Module V
Simulators, on page 69).

Effects of Friction on
Buckling
The section “General Force Balance” on page 11 clearly illustrates the role
friction plays in a force balance and the transmission of force along the CT
string, especially after the CT is helically buckled. However, the discussion
to this point has ignored effects of friction on the onset of buckling.

26 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

MEPTEC conducted detailed experimental buckling studies with small-


scale laboratory equipment and actual CT inside 400 ft of casing to
determine what factors affect the buckling and post-buckling behavior of
tubulars. Results from MEPTEC's buckling experiments indicate that
friction significantly affects buckling behavior of rods and tubing.

! NOTE: In general, friction stabilizes a tubular under compression to delay


the onset of buckling. Friction also causes hysteresis in the post-buckling
behavior. As a result of this hysteresis, buckling behavior is different from
unbuckling behavior.

Figure 14 shows typical results from several laboratory experiments. The


data are for two different aluminum rods subjected to buckling inside a
horizontal acrylic tube. One end of each rod was fixed, while a remote-
controlled linear actuator applied axial force on the other end. Figure 14
shows the behavior of the axial force at the stationary end of the rod as a
function of axial displacement of the other end. The experiments actually
continued beyond 0.20 in displacement, but the plot was truncated for
simplicity.

! NOTE: The figure clearly shows that hysteresis is significant, and the axial
compressive force at unbuckling is always lower than axial compressive
force at onset of buckling.

MEPTEC obtained similar results from the large-scale tests with CT


subjected to buckling inside casing.

July 30, 1996 Proprietary information of Mobil 27


Module II Mechanical Performance Mobil Coiled
Tubing Manual

Onset of helical buckling


25 - during loading

20 _ Onset of sinusoidal buckling


_ during loading

End of helical buckling


during unloading

End of sinusoidal buckling


during unloading

0.10 0.15
Free End Displacement (in)

Figure 14:
Effects offiiction on buckling of rods

b@ NOTE: Comparison of these results with theoretical predictions shows that


current theory actually predicts axial compressive force at
unbuckling.

As a result, expressions such as Eq. 20 and Eq. 21 predict critical buckling


forces significantly lower than those determined experimentally. Since
friction is always present in the real world, predicted critical buckling
forces are conservative. Unfortunately, no theoretical work has studied
frictional effects. This is partly due to the difficulty of including friction in
traditional analysis methods for buckling studies.

Friction has a stabilizing effect on buckling, i.e., it delays the onset of


buckling compared to the frictionless case. One way of modeling this
effect is to account for the additional drag force a segment must overcome
to buckle. Based on the force balance derived in the General Force
Balance section (page 1 l), drag on a segment of unit length is
Cf x W sin 6. Adding this frictional stabilizing force to the normal force in

28 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

the adjusted critical helical buckling limit for an inclined segment with
friction in Eq. 23.

Equation 23: Adjusted critical helical buckling limit for an inclined segment with
friction

EI × W sinθ
′ = 2 2(1 + Cf )
FCH = ( )
1 + Cf FCH (23)
Ø rc

where the critical helical unbuckling limit, FCH , is given by Eq. 21.

Equation 23 provides a simple method to account for frictional effects that


agrees better with experimental results than Eq. 21. However, it still seems
to underestimate the measured critical helical buckling force. Further
theoretical and experimental work could improve the ability to model
effects of friction on buckling. Until then, use Eq. 23 to account for
frictional effects on buckling.
Equation 23 implies that higher values of Cf are beneficial for CT
operations. Higher friction delays buckling, which results in lower drag and
longer reach for RIH compared to buckled tubing. However, once the CT
buckles, higher Cf means post-buckling drag will be significantly higher
than drag for unbuckled CT. Equation 19 through Eq. 23 predict buckling
only for a single CT segment, i.e., locally. A microscopic view gives no
indication of buckling elsewhere in the CT string. Any increase in drag on a
segment of CT increases axial compressive force on segments above it.
Thus, increasing Cf at one location may simply shift the buckling problem
up the wellbore. (see Module V, Simulators, on page 69)

! NOTE: A good CT simulator is required to model the macroscopic effects


of friction.

Post-Buckling Lock-up
By itself, helical buckling is neither a critical problem nor a limiting
condition for CT.

! NOTE: Aside from exceeding mechanical limits of the tubing, the limiting
condition for CT operations during RIH is post-buckling lock-up.
However, for horizontal CT with no BHA (Fb=0), it is possible to estimate
the maximum reach prior to lock-up using Eq. 24.

Lock-up not only limits SDW or WOB, it can also prevent the BHA from
reaching TD. In simple terms, lock-up is a local phenomenon that occurs

July 30, 1996 Proprietary information of Mobil 29


Module II Mechanical Performance Mobil Coiled
Tubing Manual

during RIH when the increase in drag exceeds the increase in axial
compressive force. When buckled CT reaches this condition, any further
increase in axial compressive force is lost completely to drag. Since normal
force due to helical buckling increases as the square of axial compressive
force, Eq. 9, lock-up may occur almost immediately after a segment
helically buckles. Attempting to force more CT into a hole after lock-up
can damage the tubing.

! NOTE: Proper modeling of post-buckling drag effects with a good CT


simulator is necessary to determine whether lock-up occurs.

Equation 24: Approximate maximum reach prior to lockup for horizontal CT


with no BHA

π  FCS   π  ′
FCH
HL = ×  ×  − tan −1 π (1 + Cf )  +
2  (W + Cf )   2  (W + Cf )
Ø
(24)

! NOTE: Use Eq. 24 to eliminate from further consideration any CT that


would lockup in the horizontal section before reaching TD.

Example: Washing sand out of a horizontal well

It is desired to use CT to wash sand out of the 5600-ft horizontal section of


a well. The horizontal section was completed with 5.50 in x 23.0 lbs/ft
casing, IDhole = 4.670 in. A reel of tapered 1.50-in CT is available with
6500 ft of 0.125-in wall at the bottom. Will this CT string be suitable for
the job?

Since the well is full of water (ρo = 8.4 lbs/gal), a reasonable value of Cf is
0.30. Filling the CT with nitrogen (ρi = 2.0 lbs/gal), because that will help
extend its reach. For these conditions, Eq. 24 gives HL = 6020 ft, and the
existing CT string should reach the bottom of the horizontal section.
However, a CT simulator is the only way to predict the overall
performance of the CT during RIH and POOH.

30 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tu binq Manual

Mechanical
Limits
Figure 15 depicts a typical segment of CT subjected to axial, bending,
torsional, and pressure forces. These forces cause stresses, o, in the
segment that cannot exceed certain values without damaging the CT. The
development of stress theory and equations to calculate stresses in CT is
beyond the scope of this manualx~9. However, the following section
introduces the concepts and general equations for determining mechanical
limits. The Excel spreadsheet “CT Forces & Stresses” (see page 55)
contains equations to calculate stresses and limiting conditions for CT
segments.

Figure 15:
Stresses on a segment of CT

Axial and Bending


Stresses

The axial stress at any location in the segment is the axial force divided by
the cross-sectional area of the steel, as shown in Eq. 25.

* Timoshenko, S., Strength of Materials, Part II, D. Van Nostrand Co., Inc., 1956.

9 Mobil LRFD Casing Design Manual.

July 30, 1996 Proprietary information of Mobil 31


Module II Mechanical Performance Mobil Coiled
Tubing Manual

Equation 25: Axial stress at any location in the segment


Faxial (30)
σF =
Ø Asteel
(25)

where Eq. 26 gives the cross sectional area of steel in the wall of the CT.

Equation 26: Cross sectional area of steel

π[( OD 2 − (OD − 2t ) 2 ]
Asteel =
Ø 4
(26)

Axial stresses have the same sign as axial forces, positive (+) for tension
and negative (–) for compression.
A bending moment MR in a segment with radius of curvature R generates
bending stress as shown in Eq. 27.

Equation 27: Bending stress in a segment due to a bending moment


M R × OD
σ bend = ±
Ø 2I
(27)

where Eq. 28 gives the bending moment for pure bending.

Equation 28: Bending moment in a segment


EI
MR =
Ø R
(28)

Combining Eq. 28 with Eq. 27, yields Eq. 29.

Equation 29: Bending stress in a segment


E × OD
σ bend = ±
Ø 2R
(29)

The above equations are valid only as long as the bending stress does not
exceed the yield strength of the material (i.e., only in the elastic regime).
Since the bending stress is also axial in nature, total axial stress at any
location is the sum of the stresses due to axial force and bending.

Equation 30: Total axial stress at any location

32 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

Ø σ axial = σ F ± σ bend

In practice, we use Eq. 30 to evaluate axial stress at both the inner and
outer radii, with both the positive and negative sign for bending stress. This
results in four values for the axial stress. We then select the axial stresses
with the largest absolute value to include in further calculations.
The radius of curvature of a helically buckled segment, Rh is usually much
smaller than the radius of curvature R of the hole. Thus, bending stresses in
a helically buckled segment are greater than those in an unbuckled segment
at the same location. Rh depends upon the OD of the CT, bending stiffness,
radial clearance, and the local pitch of the helix. Equation 29 applies for
buckled segments with Rh in place of R. Equation 31 gives the bending
stress for a helically buckled segment with an axial force Faxial acting on
it10.

Equation 31: Bending stress for a helically buckled segment with a force
rc × Faxial × OD
σ bend = ±
Ø 4I
(31)
Torsional Stress
Equation 32 gives the torsional stress at any location due to torque T.

Equation 32: Torsional stress at any location


T × OD
τ=
Ø 2J
(32)
where J, the polar moment of inertia equals 2 x I. Torsional stress is always
positive, unlike bending and axial stresses.

Pressure Stresses
A pressure differential across the wall of the segment creates radial and
tangential stresses, Figure 15, that vary with radial position. Equation 33
gives the radial stress at any radial location r in the wall of the segment.

Equation 33: Radial stress at a location in the wall of a segment.

10
Mitchell, R.F., 1986, “Simple Frictional Analysis of Helical Buckling of Tubing,”
SPE Drilling Engineering, December 1986, pp. 303-310.

July 30, 1996 Proprietary information of Mobil 33


Module II Mechanical Performance Mobil Coiled
Tubing Manual

( Po − Pi )ri 2 ro2 ri 2 Pi − ro2 Po


σ radial = +
Ø r 2 (ro2 − ri 2 ) (ro2 − ri 2 ) (37)
(33)

where the subscripts i and o refer inner and outer radii of the CT,
respectively.

Equation 34: Inner radius of CT


OD − 2t
ri =
Ø 2
(34)

Equation 35: Outer radius of CT


OD
ro =
Ø 2
(35)
From Eq. 33, at the inner radius σ radial = − Pi , while at the outer radius
σ radial = − Po .

Equation 36 gives the hoop (or tangential) stress at any radial location r in
the wall of the segment.

Equation 36: Tangential stress at a location in the wall of a segment

ri 2 Pi − ro2 Po ( Po − Pi )ri 2 ro2


σ hoop = − 2 2
Ø (ro2 − ri 2 ) r (ro − ri 2 )
(36)

Equation 33 and Eq. 36 are known as Lame's equations11.

Total Equivalent (von


Mises) Stress
In order to assess the safety of a segment, a failure criterion is necessary.
MEPTEC uses the von Mises overall stress failure criterion to calculate
total equivalent stress on a segment. Equation 37 gives the von Mises
overall stress (sometimes called the von Mises equivalent uniaxial stress, or
the VME stess).

Equation 37: Von Mises equivalent stress

11
Timoshenko, S., Strength of Materials, Part II, D. Van Nostrand Co., Inc. 1956.

34 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

Equation 37: Von kfises equivalent stress

2 + ajl~~p2+ or;adial 2 - oaxial ohoop - aaxial orradial - oradial ohoop + 3($ + z; + +q (37)

Shear stresses, zt-~3, are normally zero for CT applications. If oVME> ovP
(the yield stress) at any location in the segment, the CT can fail.

/ NOTE: To provide a margin of safety, a Yield Factor of 0.8 is usually


applied which ensures that oME < 0.80~~. Nominal yield stresses for
common CT materials are 70- 100 ksi. Thus, “safe” oVMEfor these
materials are 56-80 ksi.

Tubing Burst Pressure

To calculate the true burst capacity of a segment of CT, assume the


following:

1. The CT material has a true elastic-plastic stress strain relationship as


depicted in Figure 16.

2. The segment wall plasticizes completely prior to burst.

3. The CT is axially restrained or capped at the free end. (This may not
apply during pumping operations while the CT is moving or suspended
off-bottom.)

Yield
Point

Figure 16:
AT Elastic-Plastic stress strain curve

Strain

With these assumptions, Eq. 38 gives the true burst capacity of a segment
of CT.

July 30, 1996 Proprietary information of Mobil 35


Module II Mechanical Performance Mobil Coiled
Tubing Manual

Equation 38: True burst capacity of a segment

 
2σ  1 
Pburst = YP ln  (38)
3  2t 
 1 − OD 
Ø
where ln denotes the natural logarithm.
Table 2 shows burst pressures calculated from Eq. 38 for common CT
sizes and σ YP = 80 ksi . Appendix G contains additional burst pressure
calculations.

Table 2: CT burst pressures (psi) for σ YP = 80 ksi

CT Wall Thickness, t (in)


OD (in) 0.095 0.102 0.109 0.125 0.134 0.156 0.188

1.500 12511 13504 14507 16842 18182 21542


1.750 12290 14240 15355 18139
2.000 10661 12335 13290 15667 19238
2.375 8894 10275 11060 13010 15921
2.875 8404 9039 10612 12948
3.500 7359 8625 10498

Tubing Collapse
Collapse is a difficult failure mode to predict accurately because it depends
on factors that are seldom known accurately. These factors include tubing
ovality, yield stress, and wall thickness. If σVME > σ YP , CT can collapse
whenever Po > Pi. Maximum tension in a CT string occurs just below the
injector head at the start of POOH, but maximum axial stress may occur
elsewhere. Also, CT may collapse due to hydrostatic pressure of fluids in
the wellbore annulus if the annulus fluid is denser than the fluid in the CT.

! NOTE: Always calculate collapse pressure for each section in a tapered CT


string using the maximum axial stress expected in that section.

Collapse is simpler to prevent than to predict. The most direct method is to


maintain Pi > Po. However, this cure may lead to another, equally serious
problem. High internal pressure in CT moving through the injector head,

36 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

across the gooseneck, and on/off the reel leads to rapid tubing fatigue (see
page 43). Consequently, preventing CT collapse can be a tradeoff against
minimizing fatigue damage.
Equation 39, first derived Timoshenko, predicts collapse pressure
differential based on triaxial stress conditions12. Equation 39 incorporates
standard geometric and material parameters, ovality, and axial stress. It
also accounts for both yield-induced collapse and stability-induced
collapse. However, it excludes effects of eccentricity and residual stresses.

Equation 39: Collapse pressure differential based on triaxial stress conditions


2
 3ϕOD    3ϕOD  
Py + Pe  1 +  −  Py + Pe  1 +   − 4 Py Pe
 2t    2t  
Po − Pi = (39)
Ø 2

where:

Equation 40: Py (yield collapse term) for collapse pressure calculation

 2t 
Py = Γσ YP  
Ø  OD 
(40)

Equation 41: Pe (stability-induced collapse term) for collapse pressure


calculation
3
2E  t 
Pe =  
Ø 1 − µ 2  OD 
(41)

Equation 42: Γ (axial stress correction factor) for collapse pressure calculation
2
σ  σ 
Γ = 1 − 3 axial  −  axial 
 2σ YP   2σ YP 
Ø (42)

12
Allen, H.G. and Bulson, P.S., “Background to Buckling”, McGraw Hill, NY, 1980.

July 30, 1996 Proprietary information of Mobil 37


IModule II Mechanical Performance Mobil Coiled
Tubing Manual

Poisson’s ratio, p, is 0.33 for steel. Equation 43 is the definition of ovality,


with 9 > 0.0 meaning oval or deformed tubing.

Equation 43: Ova&y

2(Wl, - ODmin) (43)


L ’ = OD,,,,, + OD,,,,,

The Excel spreadsheet “CT Forces & Stresses” (see page 55) includes Eq.
39 through Eq. 43 for calculating collapse pressure of CT segments.
I

/ NOTE: CT collapse pressure is highly sensitive to ovality.

Figure 17 illustrates this sensitivity for a relatively small range of ovality


with typical 1.50-in CT having a,, = 80 ksi .

Collapse
Pressure (psi)

Ovality (in/in)

Axial Load (Ibs)

Figure 17:
Calculated collapse pressure for 1.50-in OD x 0.102-h wall CT

38 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

Unfortunately, ovality is an uncontrollable factor during CT operations


that worsens with time. Newly-milled CT is round but becomes slightly
oval, around 9 = 0.005, when coiled onto a shipping reel. Each round trip
from/to a reel increases the ovality.
I I

/ NOTE: Always include ovality in CT collapse pressure calculations.

Values between 0.01 and 0.05 are reasonable for planning purposes. The
best approach is to measure the actual ovality of the CT, especially if
maximum collapse performance is critical. Some CT service companies
provide real-time measurements of tubing ovality and wall thickness. If
collapsing the CT is the most likely failure mode during an operation, such
real-time measurements could be the key to success.

Figure 18 shows the effect of wall thickness on collapse pressure


calculated from Eq. 39 for typical 1.50-in CT with a fixed tensile load.
Note the compounding effect of ovality. This figure supports the warning
at the beginning of this section, “. . . always calculate collapse pressure
for each section in a tapered CT string . . . ‘I.

Collapse
Pressure (psi)

(in/in)

Wall Thickness d

(in)

July 30, 1996 39


Module II Mechanical Performance Mobil Coiled
Tubing Manual

Uncertainties in tubing ovality, yield stress, and wall thickness, can lead to

measured collapse pressures. Unfortunately, little data has been published


for CT collapse, and they are not satisfactory for “calibrating” Eq. 39.

with OD/t and yield stress in the same range as commonly available CT. No
similar body of data is available for CT

18000

16000

14000
Collapse Pressure (psi)

12000

10000

8000

6000

4000

10.0 12.0 14.0 16.0 18.0 20.0 22.0 24.0

OD/t

Figure 19
Measured collapse pressure of casing and tubing with ovality
70 ksi < σ YP < 120 ksi

Using Eq. 39 to calculate collapse pressure for each test summarized in


Figure 19 results in the ratio of calculated and measured values in Figure
20. Based on these 232 results, Eq. 39 underestimates actual collapse
pressure 3-4%, on average. However, the scatter in collapse pressure ratio
is quite high. Figure 20 shows the range needed for 95% confidence (±2
standard deviations) in the calculated value.

40 July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

1.300

1.200 +2 sigma = 1.1487

1.100

1.000

mean = 0.9671
0.900

0.800
-2 sigma = 0.7856
0.700

0.600
8.0 12.0 16.0 20.0 24.0

OD/t

Figure 20:
Collapse pressure ratio for materials with 70 ksi< σ YP <120 ksi

The data in Figure 20 presented as a histogram in Figure 21 shows the


distribution of collapse pressure ratio.

July 30, 1996 Proprietary information of Mobil 41


Module II Mechanical Performance Mobil Coiled
Tubing Manual

25

20

15

10

0.70 0.74 0.78 0.82 0.86 0.90 0.94 0.98 1.02 1.06 1.10 1.14 1.18 1.22
Calculated/Measured

Figure 21:
Distribution of collapse pressure ratio - geometry 8 < OD/t < 24 and materials
with 70 ksi < σ YP < 120 ksi

Figure 21 indicates that multiplying results from Eq. 39 by 0.74 will reduce
the risk of collapse failure to nearly zero. Minimum Oval Collapse on the
spreadsheet "CT Forces and Stresses" includes this factor of 0.74 (see page
55). Maximum collapse resistance occurs for zero axial stress and round
tubing. Table 3 shows examples of oval collapse pressure calculated from
Eq. 39 for common CT sizes. Appendix G contains additional collapse
pressure calculations.

42 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

Table 3: Oval CT collapse pressures (psi) without safety factor

CT Wall Thickness, t (in)


OD (in) 0.095 0.102 0.109 0.125 0.134 0.156 0.188

1.500 7193 8028 8849 10684 11697 14134


1.750 7004 8632 9526 11665
2.000 5577 7043 7850 9773 12477
2.375 3985 5230 5932 7615 9971
2.875 3547 4116 5533 7563
3.500 2650 3744 5431

ϕ = 0.005 σ axial = 0 psi σ YP = 80 ksi

Fatigue Life
A tubular subjected to repeated loading and unloading or to reversal of
stresses, loses some of its resistance to failure with each stress cycle. This
phenomenon is known as fatigue. The number of stress cycles a tubular
can withstand before failing is called its fatigue life. Each stress reversal in
a tubular consumes some of its available fatigue life and adds to its
"accumulated fatigue damage". Estimating the expected fatigue life or
cycles remaining to failure for a CT string is an important part of designing
a CT operation.
For a large number of mechanical design problems, the magnitude of cyclic
stress in components subject to fatigue is within the elastic limit. In such
cases, engineers estimate the fatigue life of a component from actual
fatigue tests on the material in the component. The results from such tests
relate the magnitude of reversed stress to the number of cycles to failure at
that stress for the subject material. Thus, given the magnitude of stress
reversal, an engineer can estimate the number of cycles to failure at that
stress. Since the number of cycles to failure for elastic deformations is
extremely high, such fatigue is commonly referred to as High Cycle
Fatigue.
Sometimes, however, the reversed stresses are beyond the elastic limit, and
a component experiences reversing plastic deformations. Plastic fatigue
is a complex and poorly understood problem. However, one fact is
perfectly clear.

July 30, 1996 Proprietary information of Mobil 43


Module II Mechanical Performance Mobil Coiled
Tubing Manual

! NOTE: A component subjected to reversing plastic deformations fails after


relatively few cycles.

Unfortunately, this fatigue failure regime, Low Cycle Fatigue, applies to


CT.
CT begins its life plastically deformed, because it is wound on a reel.
Moreover, each round trip into the well and back plastically deforms the
tubing six times. These deformations are:
1. RIH - unwind from the reel
2. RIH - across the gooseneck
3. RIH - through the injector
4. POOH - through the injector
5. POOH - across the gooseneck
6. POOH - wind back onto the reel

! NOTE: During each round trip, CT plastically deforms and then straightens
three times. Each reversed plastic deformation shortens the life of the
tubing, and internal pressure amplifies the damage.

Consequently, the number of deformation cycles, extent of deformation at


each cycle, and the pressure at each cycle are required to estimate
accumulated fatigue damage for a segment of tubing. Figure 22 depicts the
variability of CT fatigue damage for numerous round trips in a well. The
larger "slices" of damage correspond to higher pressure, higher stress, or a
combination of both. The accumulated fatigue damage is the sum of the
damage for round trips 1-14.

44 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

CT Fatigue Per Round Trip

1 2
3
Remaining 4
Fatigue Life 5
6
7
14 8
13 9
12 11 10

Figure 22:
Consumption of CT fatigue life

Early attempts to predict and combat CT fatigue were strictly based on


experience, i.e., the number of round trips CT could make without
breaking. Predictions based on such historical data are not very reliable
since this method ignores the contribution of pressure to fatigue.
Moreover, these data cannot localize the CT segment(s) most likely to fail.
With the advent of better data acquisition systems, service companies can
track the number of trips for individual segments of a CT string. However,
despite access to computer data acquisition systems capable of monitoring
all aspects of a CT operation, some service companies still simply count
round trips or "running feet". These companies use hefty safety margins to
reduce their risk or add arbitrary damage multipliers to account for high
pressures, corrosion, and CT contact with acids.

! NOTE: Running feet methods are unsatisfactory because they miss the
connection between cyclic strain, pressure, and fatigue failures.

In 1990, Dowell conducted controlled fatigue tests on numerous CT


samples using a full-scale CT unit. These were the first "scientific" tests on
CT fatigue to appear in the open literature. Their objectives were two-fold:

July 30, 1996 Proprietary information of Mobil 45


Module II Mechanical Performance Mobil Coiled
Tubing Manual

1. Quantify effects of CT dimensions, gooseneck radius, and pressure


2. Provide data for validating their fledgling fatigue model, CoilLIFE
In 1993, MEPTEC participated in a joint industry project (JIP) coordinated
by Dowell that:
• Designed, built, and operated a fatigue test machine for CT (1.25-3.5 in
OD)
• Quantified effects of numerous factors on CT fatigue life

• Improved and validated the CoilLIFE program for predicting CT fatigue

Appendix C provides detailed information on this JIP. In summary, results


from the fatigue test machine compare favorably with Dowell's full-scale
tests, and the CoilLIFE program gives reasonably accurate estimates for
CT fatigue life. From a practical standpoint, the most important product of
this JIP is the relationship between measurable parameters and CT fatigue.
About the same time, NOWSCO privately designed, built, and operated a
CT fatigue test machine and started developing their own CT fatigue
model. Nowsco’s experimental and theoretical work culminated in their
proprietary fatigue model, CYCLE. CYCLE also does reasonably well in
comparison with experiments and includes some reliability-based concepts
as well.

! NOTE: For a given material, CT fatigue life increases with:


• Increasing tubing wall thickness
• Decreasing OD
• Decreasing internal pressure
• Increasing gooseneck radius
• Increasing reel diameter

Of these parameters, internal pressure is the only parameter under


operational control since job requirements determine the others. CT with
internal pressure balloons slightly (grows diametrically) during each plastic
deformation. Obviously, ballooning decreases CT fatigue life by thinning
the tubing wall and increasing its OD. Some service companies
continuously monitor tubing OD to ensure that the tubing will pass through
the stripper elements. A typical limit for diametrical growth is 6%.
However, this limit does not protect the tubing against fatigue failure,
because ballooning does not directly correlate with fatigue life.

! NOTE: Even with the best of fatigue models, predicting remaining fatigue
life for a used reel of CT is impossible without its complete operational
history. Thus, service companies should maintain a complete history of

46 Proprietary information of Mobil July 30, 1996


Mobil Coiled Module II Mechanical Performance
Tubing Manual

each CT reel in their inventory. If they do not, Mobil cannot depend on


their fatigue life predictions. Critical CT jobs may require a new reel of CT.

Measuring CT fatigue life requires destructive test(s) that render the tubing
useless. Conducting such tests on CT segments removed from the free end
of a reel is a waste of time since the results indicate nothing about the
fatigue life of segments far up the reel. However, a CT operator can take
several practical steps to minimize the risk of fatigue failure including:
• Reducing CT internal pressure during trips

• Avoiding short-cycles

• Using the largest reel diameter available

• Designing CT strings with the thickest wall possible

• Using the largest gooseneck radius possible

• Using the highest yield strength material possible

Any operation that repeatedly cycles a given segment of tubing over the
gooseneck (short-cycling) significantly increases the risk of fatigue failure
in that segment. Such an operation concentrates fatigue damage over a
relatively short length of tubing. Pumping through the CT during short-
cycling compounds this fatigue damage. Therefore, regularly cutting a
length of tubing from the free end greatly prolongs its useful life by moving
the most heavily damaged CT segment(s) away from the source of the
damage (reel, gooseneck, and injector head).

July 30, 1996 Proprietary information of Mobil 47

Das könnte Ihnen auch gefallen