Sie sind auf Seite 1von 16

Journal of Petroleum Science and Engineering 150 (2017) 272287

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


journal homepage: www.elsevier.com/locate/petrol

Dynamic model for longitudinal and torsional motions of a horizontal MARK


oilwell drillstring with wellbore stick-slip friction

Mejbahul Sarkera, , D. Geo Rideoutb, Stephen D. Buttc
a
Oil and Gas Engineering, Memorial University of Newfoundland, Canada
b
Department of Mechanical Engineering, Memorial University of Newfoundland, Canada
c
Department of Process Engineering, Memorial University of Newfoundland, Canada

A R T I C L E I N F O A BS T RAC T

Keywords: Horizontal oilwell drillstring failures are very costly. Vibration causes excessive wear and tear on the bottom
Horizontal drilling hole assembly and reduces the life of the drill bit, and has motivated extensive research on these types of
Wellbore friction drillstring vibrations. Two key factors in simulation of motions of horizontal drillstrings are to have a model
Vibration capturing coupling among various types of vibration, and an accurate but ecient treatment of the wellbore
Rate of penetration
friction. In this paper, a bond graph dynamic model of a horizontal oilwell drillstring has been developed that
Lumped segment approach
predicts longitudinal motion, torsional motion, and coupling between longitudinal and torsional motion excited
Finite element
Bit-rock interaction by bit-rock interaction. A lumped segment modeling approach of vertical drillstring dynamics has been
Agitator tool extended to include dynamic wellbore friction in the build and horizontal sections of the drillstring. The model
Simulation incorporates torsional viscous damping, longitudinal hydrodynamic damping, and buoyancy eect due to
drilling uid; an extended bit-rock interaction model that allows the drillstring to advance at the rate of
penetration, a downhole mud motor, and top drive ac motor dynamics. The friction coecient between
drillstring and wellbore has been tuned with the aid of eld data from a horizontal oileld in Canada. The model
predicts the expected coupling between weight on bit, bit speed, and bit-rock interaction conditions; and their
eect on longitudinal and torsional motions. Finally, an experiment was conducted with a downhole axial
vibration tool (Agitator). A force excitation source, which simulates the Agitator tool, in the longitudinal
direction has been implemented in the horizontal section of the virtual drillstring. Simulations show a better
weight transfer to the bit, with a low frequency and high amplitude force excitation giving best performance.

1. Introduction shown in Fig. 1. MWD tools sometimes fail due to excessive vibrations.
Their high cost along with cost of tripping to replace such tools,
Horizontal oilwell drilling is, necessary to maximize production motivates the development of sophisticated drillstring vibration models
from a formation and reduce environmental impact at the surface. to understand and prevent conditions that lead to failures. One of the
Vibrations are always present to some degree while drilling in deviated key engineering challenges related to building horizontal drillstring
wells but can be especially bad in dicult drilling environments (e.g. dynamic models is the modeling of wellbore friction. While drilling
hard formations, steep angle wells). Vibration can aect weight-on-bit horizontal wells, especially extended-reach wells, friction between the
(WOB), rate-of-penetration (ROP), and drilling direction and can also drillstring and borehole wall is the main source of decreasing ROP
severely damage the bottom-hole-assembly (BHA), measuring-while- caused by energy lost due to poor WOB transfer. One of the rst
drilling (MWD) tools and drill bit cutters. Field experience suggests contributions to the work of understanding friction in the well has been
that drillstring vibrations and related failures can account for approxi- presented in (Johancsik et al., 1984) developing a torque and drag
mately 210% of well costs (Jardine et al., 1994). Since horizontal model with basic equations for friction in deviated wellbores. It was
drilling is one of the most expensive operations in oil exploration and assumed that both torque and drag are caused entirely by sliding
development, drilling with optimum operational parameters such as friction forces that result from contact of the drillstring with the
WOB, top drive rotational speed, downhole mud motor speed, torque- wellbore. Later the work was put in a standard dierential form in
on-bit (TOB) and bit hydraulic horsepower is required from an (Sheppard et al., 1987). A larger study on friction analysis for long
economic point of view. A schematic of horizontal oilwell drilling is reach wells has been undertaken in (Aadnoy, 1998). All models


Corresponding author.
E-mail addresses: mms426@mun.ca (M. Sarker), g.rideout@mun.ca (D.G. Rideout), sdbutt@mun.ca (S.D. Butt).

http://dx.doi.org/10.1016/j.petrol.2016.12.010
Received 13 September 2016; Received in revised form 29 November 2016; Accepted 5 December 2016
Available online 10 December 2016
0920-4105/ 2016 Elsevier B.V. All rights reserved.
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

which friction is function of velocity. The model has the advantage of


generating ordinary dierential equations but can still experience
numerical instabilities in the stick phase. A switch model proposed in
(Leine, 2000) consists of three dierent sets of ordinary dierential
equations for the stick, slip and the transition phases. At each time step
the state vector is inspected to determine whether the system is in the
slip mode, in the stick mode or the transition mode. The corresponding
time derivative of the state vector is then chosen. A region of small
velocity is dened for the stick band and the system is considered to be
in the slip mode if the relative velocity lies outside this narrow stick
band. In one state the velocity is prescribed and the force is
determined, and in other state, the force is prescribed and the velocity
is determined. Such causal inversions create formulation and compu-
tational problems, and these problems can be quite prohibitive if many
switches are part of the model.
A modication of the Karnopp's model is presented by Margolis
(2004) that allows the stick-slip friction element to be self-contained,
which is represented as a combination of dissipative and elastic
elements in a bond graph. The elements require a velocity input from
the attached system and output the friction force similarly to Karnopp's
model. The dierence can be identied during the stuck phase where
the friction force continues to be calculated internally to the element
and does not require any information from the attached system. The
model is self-contained because the tests of the stuck and unstuck
states have no dependence on the overall systems to which the friction
Fig. 1. Sketch of horizontal oilwell drilling system (courtesy of Schlumberger). generated elements are attached.
The stick-slip friction model proposed in this paper takes a similar
discussed above are soft-string models and the stick-slip phenomena in modeling approach to Margolis. A bond graph C-element (compliance
the friction model has been ignored. A computationally ecient yet with some logical modication of build-in codes) simulates the stick-
predictive wellbore friction model remains an open research problem. slip phenomena. The output of the C-element is the friction force. The
To the best of the author's knowledge, no complete dynamic model for input velocity, which is the relative motion between the contact
a horizontal oilwell drillstring has been developed. surfaces, allows determination of the stuck and unstuck states. The
This paper is organized as follows. Section 2 develops, implements necessary logical information is shown in Figs. 2 and 3. In the stick
and validates a stick-slip friction model that will allow the simulation to phase, the friction force is generated by the small but nite deformation
accurately capture an important source of energy loss during drilling of a high sti spring-damper system which represents deformation of
and tripping. In Section 3, dynamic normal forces for the friction contact surface asperities. When the force exerted by the spring-
submodel are calculated based on a modication of an existing static damper on the system mass becomes equal to the maximum static
treatment. A lumped-segment drillstring model, with coupled axial and friction force and the relative motion is still in the stick band, then the
torsional vibratory motions, is presented in Section 4, followed by a bit- spring state (deection) is set at a constant value in order to create a
rock interaction model in Section 5 that allows the drillstring and bit to constant static friction force output. During the slip phase the output
advance in the borehole. Field data is used in Section 6 to tune the from the C-element is simply the kinetic friction force. The model from
friction factor. Given that the authors and their industry partners are Figs. 2 and 3 are simulated in 20Sim bond graph software using the
motivated to predict the vibrations on and induced by downhole tools, proposed self-contained friction model. The bond graph model of the
Section 7 summarizes an experimental program by which an axially- mass-surface system is shown in Fig. 4. As described in Appx. A,
vibrating tool (Agitator) was characterized for use in the simulation. In elements bonded to the 1-junction have a common generalized ow
Section 8, the complete horizontal drillstring model is used to show the (velocity) and their generalized eorts (forces) sum to zero. Therefore,
eect of downhole tool output on WOB, ROP, and vibration levels at
multiple locations. The model is a potentially valuable tool in the
design of drillstrings with optimized top drive speeds, stabilizer and
downhole tool locations, mud motor speeds, and trajectories. The
model development was facilitated by the bond graph approach, an
overview of which is given in Appx. A.

2. Modeling of stick-slip friction phenomena

The stick-slip nature of friction is very common when the relative


velocity between sliding surfaces approaches zero and the surfaces
become stuck, requiring a force larger than the sliding friction force to
break the surface loose. The most basic friction models contain
Coulomb friction and linear viscous damping which describe the
friction forces well for steady state velocities. When velocity crosses
zero most models present numerical problems. To overcome these
problems during simulations, Karnopp (1985) proposed a friction
model to set the friction force equal to the external forces acting on
the object, for a small neighborhood around zero velocity, outside of Fig. 2. Physical schematic of stick-phase.

273
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

Table 1
Data used in mass-block system simulation.

Parameters Value

Mass, m 10 kg
ks, Cd 107 N/m, 105 N-s/m
s, k, Vthreshold 0.3, 0.2, 0.0005 m/s

Fig. 3. Physical schematic of slip-phase.

the 1-junction captures the fact that the applied force source moves
with the mass, and simultaneously enforces Newton's Second Law.
Elements bonded to the 0-junction have a common force, and their
velocities sum to zero in accordance with the power bond half-arrow
directions, which indicate the direction of algebraically positive power
ow. In Fig. 4, all elements bonded to the 0-junction are subjected to
the friction force Ff. The velocity input to the generalized compliance
(C-element), is therefore the relative velocity (Vrel) between the mass
and its sliding surface. As also described in the Appx. A, the short
strokes normal to the bond indicate the input-output structure of the
constitutive laws of the bonded elements. These causal strokes give a
visual indication of causal conicts when submodels of a complex
system such as an oilwell drillstring are assembled. Table 1 summarizes
Fig. 5. Simulation results for the mass-surface system, F(t) =35sin(50t) N.
all relevant data that is used in the mass-surface system simulation.
Fig. 5 shows the simulation results of the mass-surface system
3. Normal contact force for wellbore friction
when the amplitude and frequency of the applied sinusoidal force are
35 N and 50 rad/s, respectively. At the start of the simulation, the
Early analytical wellbore friction models discussed in (Johancsik
mass-surface is in the stick zone, the force output of the C-element is
et al., 1984; Sheppard et al., 1987; Aadnoy, 1998) are most suitable for
equal to the applied force and the C-element starts to act as an ideal
the case of drillstring pulling in/out operations, in which drillstring
compliance element with parallel damping. The initial positive slope
rotation is negligible compared to axial motion. In this paper, the
from the friction force plot is due to deformation of the spring element.
analytical model presented in (Johancsik et al., 1984), which is still
The at portion at the tip of the slope indicates that the model is
very popular in the drilling industry because of its simplicity, has been
limiting the friction force at the stick phase by constraining the
modied for drilling operations with string rotation. One of the main
deection of the C-element to be constant. The mass starts to slip
modications has been the addition of stick-slip friction phenomena
when the applied force overcomes the maximum allowable static
instead of sliding friction. In this paper, horizontal wellbore friction has
friction force and the output of the C-element becomes a constant
been included. Friction in the vertical section of the drillstring has been
kinetic friction force. The mass velocity plot shows repeating stick-slip
neglected. A sketch of the build section of the horizontal drillstring
phenomena.
segments is shown in Fig. 6. The section is divided into curve elements.
It has been assumed that the drillstring contacts at the upper face of the

Fig. 4. (a) Sketch of the system and (b) Bond graph model of the friction-element with system.

274
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

Fig. 6. Physical sketch (left) and normal contact forces (right) of drillstring segments in
the build section.

Fig. 8. (a) Physical sketch of drillstring contact with wellbore and (b) Free body diagram
of curved drillstring segment when weight dominates tension.

Fig. 7. (a) Physical sketch of drillstring contact with wellbore and (b) Free body diagram
of curved drillstring segment when tension dominates weight.

Fig. 9. (a) Physical sketch of drillstring contact with wellbore and (b) Free body diagram
wellbore when drillstring segments are under tension. A free body of curved drillstring segment when segment under compression.
diagram of a curve element is shown in Fig. 7 when the drillstring is
moving in the downward direction and the normal force can be written been assumed as equal to the buoyancy weight W of the drillstring
as below, segments.
Fn = Ft W sin (1)
4. Modeling of drillstring segment motions
where Ft = tension force acting in the curved segment
A lumped-segment approach is used in the longitudinal and
= increment of inclination angle
torsional motion models. In the lumped segment approach, the system
W = segment buoyancy weight
is divided into a series of inertias, interconnected with springs. The
= average inclination angle of the segment
accuracy of the model depends on the number of elements considered;
however, in contrast to a modal expansion approach, the analytical
There will be a neutral point where the upper portion of the
mode shapes and natural frequencies need not be determined. If a
drillstring experiences tension force and lower portion is in compres-
system model is divided into a larger number of elements, then the
sion. In the upper portion, close to the neutral point, the term Ft
accuracy of the results will be higher. The behavior will approach that
becomes less than the term W sin in Eq. (1). The situation is depicted
of a continuous system as the number of segments approaches innity.
in Fig. 8 and the drillstring segment contacts at the bottom of the
A physical schematic of the lumped-segment models is shown in
drillstring. Thus the normal contact force equation from the free body
Fig. 10.
diagram of curve segment in Fig. 8 can be written as
The model accounts for the eect of drilling uid circulation in the
Fn = W sin Ft (2) drillstring and the annular space between the drillstring and the
wellbore, on drillstring motions. The drilling uid was characterized
Finally, the normal contact force for the case of curved segment by the ow rate developed by the mud pumps. Nonlaminar Newtonian
under compressive force, which is shown in Fig. 9, can be written as ow formulations are used in calculation of uid drag force/damping
Fn = W sin + Fc (3) for the longitudinal motion. Hydrodynamic damping due to drilling
uid circulation in the drillstring and the annular space was considered
where Fc = compressive force acting in the curved segment in the longitudinal direction instead of viscous damping. In the case of
torsional motion, the viscous damping which results from the contact
3.1. Horizontal section of drillstring between drillstring surfaces and the drilling uid was considered. In
addition, the model considers the self-weight eect and buoyancy eect
In this paper, it has been assumed that the drillstring segments due to drilling uid. Figs. 1113 show the schematic of drillstring axial
contact the wellbore at the bottom side and thus the normal forces have segment model with the FBD of axial and torsional segments of

275
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

5. Bit-rock interaction motion

A prior bit-rock interaction model in (Yigit and Christoforou, 2006),


which provides coupling between longitudinal and torsional drillstring
motions, was modied in (Sarker et al., 2012) for simulating the
friction phenomena while drilling the horizontal oilwell. The model
incorporated threshold force and the eect of instantaneous WOB and
bit rotation speed on the cutting TOB. Below a threshold force, the drill
bit does not penetrate into the rock, leaving only friction as a source of
TOB. One of the major limitations in the previous bit-rock interaction
model is that the drill bit could not move longitudinally as the drill bit
cut the rock formation. Thus, the bit-rock model has been modied
Fig. 10. Physical schematic of (a) axial segments and (b) torsional segments. accordingly. This has the important benet of allowing prediction of
ROP. The dynamic WOB equation has been modied as follows:

k c (x s ROP ) if x (s + ROP )
vertical, build (curved) and horizontal sections of drillstring. The terms
WOB =
Vp and Va in the gures indicate drilling mud velocity inside the 0
if x < (s + ROP ) (5)
drillstring and the annulus, respectively. The reader is referred to
(Sarker et al., 2012) for the equations of the uid drag forces and where kc, s and ROP indicate formation contact stiness, bottom-hole
rotational uid friction to the drillstring motions. A bond graph model surface prole and rate of penetration. The physical sketch of the
for longitudinal and torsional motions of a vertical drillstring segment contact between drill bit and rock formation is shown in Fig. 17, and
is shown in Fig. 14. The axial segment bond graph shows a mass (I the bond graph model is shown in Fig. 18. The reader is referred to
element) and gravity force source (Se element) associated with segment (Sarker et al., 2012) for the equations of the bit-rock interaction model.
velocity v. Hydrodynamic dissipative forces (R elements) also con-
tribute to Newton's Second Law of the mass, with the ow sources (Sf) 6. Tuning of friction factor
and 0-junctions calculating relative uid ow velocities inside and
outside the pipe. The dissipative forces are functions of these relative Overall simulation results will be sensitive to the friction factor
velocities. Axial compliance and material damping of the segment of between drillstring and wellbore contact surfaces. Researchers from
the segment are modeled by parallel compliance (C) and dissipative both academia and industry (Aadnoy and Andersen, 2001, 2010;
elements, the forces of which are functions of the relative velocity Tveitdal, 2011) have recommended friction factors between 0.1 and
(calculated by the 0-junction) of the segment with respect to the 0.4. The friction factor should be tuned with eld data.
adjoining segment. The buoyancy weight of the drillstring segment acts
in the longitudinal direction for the case of vertical drilling. It is not the
same while drilling the build (or curve) section where a portion of 6.1. Horizontal well eld data
buoyancy weight acts in the longitudinal direction, and is shown in
Fig. 15 as an eective weight. For the case of horizontal section drilling, Friction factor is tuned using data from a horizontal well in the
there will be no contribution of buoyancy weight in the longitudinal Septimus eld in British Columbia (BC), Canada. The total measured
direction. The curve and horizontal drillstring segment models have the depth (MD) of this well is 4340 m and true vertical depth (TVD) of this
friction terms (Figs. 15 and 16), whereas friction loss has been well is 2014 m. This well has been designed to be a single build section
neglected in the vertical sections. The friction elements (C elements) after kick-o point followed by a long horizontal section. The well
in the bond graph model shown in Figs. 15 and 16 provide drag force trajectory is shown in Appx. B, Fig. B1. Sketches of drillstring
for longitudinal motion and transverse frictional force which multiplies congurations for dierent depths are shown in Fig. B2. Table B1
with drillstring radius to provide frictional torque for torsional motion. summarizes drillstring conguration as a function of length. The time-
depth plot shown in Fig. 19 shows that the drilling rate was high in the
vertical section compared to build section. This can be veried with the

Fig. 11. Schematic of (a) drillstring axial lumped segment model showing drilling uid ow and (b) FBD of axial segment and (c) torsional segment of vertical section of drillstring.

276
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

Fig. 12. Schematic of (a) drillstring axial lumped segment model showing drilling uid ow, (b) FBD of axial segment, (c) frictional torque diagram and (d) FBD of torsional segment of
build (curved) section of drillstring.

time-ROP plot shown in Fig. 20. For the horizontal section, the average From the eld results it is clear that the analysis of wellbore friction
drilling rate is better compared to the build section. The drillstring is an important factor in drilling and well design.
static weights for dierent depths are shown in Fig. 21. The static
weight decreases after passing the build section because of removing 6.2. Selection of friction factor
the heavy weight drill pipe (HWDP) from the drillstring conguration
which can be identied in the drillstring conguration chart in Appx. B. The bond graph model of the horizontal oilwell drillstring motions
Fig. 22 show the drag forces due to friction between the drillstring and has been implemented in the commercial bond graph modeling soft-
wellbore during pulling the drillstring up and pushing the drillstring ware 20Sim. With 20Sim, models can be entered as equations, block
down, respectively. Negligible drag forces are encountered in the diagrams, bond graphs, or physical components. 20Sim is widely used
vertical section and a signicant amount can be noted in the build for modeling complex multi-domain systems and for the development
section but the drag increases signicantly in the horizontal section. of control systems. The drillstring model has three main parts: long-
Also, the drag forces during pulling the drillstring up from the well are itudinal motion submodel, torsional motion submodel and bit rock
larger than the drag forces during pushing the drillstring down. submodel. The top of the drillstring is subject to a tension force (or
Figs. 23 and 24 show the surface torques which are required for hook load). Rotary motion is applied by an ac motor, through a gear
rotating the drillstring during the o-bottom and on-bottom condi- box, to the drillstring. Here an induction motor model has been used to
tions, respectively. O-bottom torque is caused by the friction between simulate the top drive motor dynamics. The reader is referred to (Rabbi
drillstring and wellbore, whereas the on-bottom torque (drilling et al., 2015) for the bond graph modeling of a three phase induction
torque) is the summation of the torque required to overcome the motor which is adopted in this paper. The lumped segment approach,
friction and the torque required to cut the rock formation. A signicant which is used in the modeling of continuous shafts, beams and rods,
amount of o-bottom torque in the vertical section indicates contacts gives the exibility to specify the segment length independently of the
between the drill sting and wellbore. The increasing o-bottom torque number of segments in the whole model. Here, a total of 10 segments
in the build and horizontal sections is due to the expected higher has been used for the vertical section. The curved portion of the
contact area between the drillstring and wellbore. The on-bottom drillstring has been divided into 20 segments, and 25 segments have
torque plot in Fig. 24 shows a constant torque in the vertical section been chosen for the horizontal portion of the drillstring. Thus the whole
which can be assumed as a torque required for cutting the rock model consists of a total of 10 segments in the simulation of vertical
formation. In the build section, the increasing on-bottom torque with section which is up to a 1720 m MD in the well chart (Appx. B) and
depth proves the importance of torque due to contact-friction and the when the drillstring exceeds the KOP and goes to the curve section then
necessary of capturing this in the simulation model. Even higher on- the whole model consists a total of 30 segments. Finally, a total of 55
bottom torque while drilling the horizontal section veried the segments have been used to simulate the complete drillstring, includ-
presence of high frictional torque due to large contact area between ing the horizontal portion. Table C1 (Appx. C) summarizes all relevant
the drillstring and wellbore. data that is used in the current simulation.

Fig. 13. Schematic of (a) drillstring axial lumped segment model showing drilling uid ow, (b) FBD of axial segment, (c) frictional torque diagram and (d) FBD of torsional segment of
horizontal section of drillstring.

277
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

Fig. 14. Bond graph segment model for (a) longitudinal (or axial) and (b) torsional motions of vertical section of drillstring.

Fig. 15. Bond graph segment model for (a) longitudinal (or axial) and (b) torsional motions of build (curved) section of drillstring.

Figs. 2527 presents the results obtained from the drillstring Fig. 22 are attributed to low resolution of the eld data logging. From a
lumped segment model. Static weights of the drillstring congurations qualitative comparison between model results and eld data, the
based on the chart in Appx. B at dierent measuring depths are shown dynamic and static friction coecients are recommended as 0.35 and
in Fig. 25. The static weight of the drillstring at 2550 m MD is lower 0. 4, respectively. Another validation of friction factor has been
than the static weight at 2160 m MD which is consistent with the eld conducted through rotating the drill string in o-bottom condition
result shown in Fig. 21. As mentioned earlier, the reason is the absence and the torques required to overcome the friction at dierent depths
of HWDPs in the drillstring conguration at 2550 m MD. The static are shown in Fig. 28. Dynamic and static frictions coecients of 0.2
weight of the drillstring is increased again in the horizontal section and 0.25 give the best match with eld data as shown in Fig. 23. Thus,
because of addition of heavy weight drill pipes. In order to tune the rotating friction coecients (static and dynamic) for drillstring rota-
friction factor between the drillstring and wellbore, the tripping in and tional motion dier from longitudinal friction coecient (static and
out operations have been conducted in the model and the results of dynamic). This assumption has a good agreement with the work
drag force required to overcome the friction are shown in Figs. 26 and presented in (Lesage et al., 1988). The negative eect of friction on
27, respectively. Discrepancies compared to the eld results shown in drilling performance can be mitigated through axially-vibrating down-

Fig. 16. Bond graph segment model for (a) longitudinal (or axial) and (b) torsional motions of horizontal section of drillstring.

278
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

Fig. 17. sketches show (a) a lobe pattern of formation surface elevation, and (b) bit and Fig. 21. Static weight vs. depth of CNRL HZ Septimus C9-21-81-19 well.
rock spring-damper representation when x < p and (c) bit contact with rock when x > = p
rock spring and damper under compression.

Fig. 22. Drag force vs. depth of CNRL HZ Septimus C9-21-81-19 well.

Fig. 18. Bond graph model of bit-rock motion.

Fig. 23. O-bottom torque vs. depth of CNRL HZ Septimus C9-21-81-19 well.

Fig. 19. Measured depth vs. drilling day of CNRL HZ Septimus C9-21-81-19 well.

Fig. 24. On- bottom torque vs. depth of CNRL HZ Septimus C9-21-81-19 well.
Fig. 20. Average ROP vs. drilling day of CNRL HZ Septimus C9-21-81-19 well.
7. Experimental characterization of downhole tool

hole tools. Such a tool will be incorporated into the simulation, and In order to reduce the friction energy loss, National Oilwell Varco
results generated in Section 8. In the next section, the experimental (NOV) has developed and manufactured the Agitator tool which is
characterization of such a tool by the author's research group is capable of producing axial oscillations down-hole. It has been proved to
presented. be an eective method to convert friction from static to dynamic and

279
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

Fig. 25. Static weight vs. depth of the model of CNRL HZ Septimus C9-21-81-19 well.

Fig. 29. Testing frame and associated pumping facility in Advanced Drilling Laboratory
at MUN.

Fig. 26. Upward motion drag force vs. depth of the model of CNRL HZ Septimus C9-21-
81-19 well.

Fig. 30. (a) Load cells, (b) upstream pressure sensor and (c) downstream pressure
sensor used in testing frame in Advanced Drilling Laboratory at MUN.

reduce the overall energy loss. A 25% friction reduction can be achieved
Fig. 27. Downward motion drag force vs. depth of the model of CNRL HZ Septimus C9-
by using the Agitator tool (Skyles and Amiraslani, 2012). To determine
21-81-19 well.
the pressure, ow, and force characteristics of the tool, a testing frame
has been built in the Advanced Drilling Laboratory (ADL) at Memorial
University of Newfoundland (MUN) that is capable of measuring
upstream and downstream pressures and resulting axial force. The
testing frame for the Agitator tool experiment is shown in Fig. 29.
Three load cells (Fig. 30), each with a capacity of 5000 lb, have been
installed underneath each corner of the triangular plate. The Agitator
tool has been installed at the middle of the triangular plate and xed
laterally from bottom to top using three sets of constraints. Sensors
(Fig. 30) record the upstream and downstream pressures. The up-
stream and downstream pressure transducers have the ranges of 0
4000 psi and 01500 psi, respectively, and temperature range from
(20) to (+80) degrees Celsius. High pressure hose has been attached
at the top of the tool that allows the tool to vibrate axially. A ow meter
has been installed at the outlet of the pipe to measure the ow rate.
Three ball valves have been put in the set up that can be operated
manually in order to isolate the Agitator unit from the main stream
Fig. 28. O-bottom torque vs. depth of the model of CNRL HZ Septimus C9-21-81-19
line. The input ow is supplied through the mobile pumping unit
well.

280
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

Fig. 33. Outlet pressure uctuation at 70 gpm ow rate.

Fig. 31. Mobile DAQ system in Advanced Drilling Laboratory at MUN.

shown in Fig. 29. The unit can deliver a ow rate up to 70 gallons per
minute (gpm) with maximum pressure of 2500 psi. A sophisticated 16-
channel portable data acquisition (DAQ) system shown in Fig. 31,
designed to work in harsh environments, has been used to read the
data from sensors. The DAQ has a NI9188 chassis and a NI9237 for
Fig. 34. Axial force prole generated from Agitator tool at 70 gpm ow rate.
acquiring the data.
A bypass valve, which isolates the Agitator tool from any ow of
water by direct connecting the inlet and outlet ow line, is used to set
the desire ow rate. Once the desired ow rate is set, the inlet and
outlet valves for the tool is opened and the bypass valve is closed to
allow the water go across the tool. The data is recorded for about 20 s
after 1 min of the tool operating time. Once the enough data obtained
the recording is stopped and the bypass ow line is opened to isolate
the tool again from any ow. The above procedure repeated for a
dierent set of ow rate.
Figs. 3237 show the experimental results found from the Agitator
tool. All tests have been done at atmospheric pressure. The inlet and
outlet pressure uctuations at 70 gpm ow rate are shown in Figs. 32
and 33, respectively. A pressure uctuation of 570 psi was measured.
The generated force from the Agitator tool at 70 gpm ow rate is Fig. 35. Spectrum of tool generated force at 70 gpm ow rate.

shown in Fig. 34. An oscillation of 400 lbs was observed. The dominant
frequency was found to be 20 Hz as seen in frequency spectrum of the
force oscillation shown in Fig. 35. The oscillating frequency increases
with ow rate. The plot of oscillation frequencies at dierent ow rates
is shown in Fig. 36. The pressure drop across the agitator tool also
depends on the ow rate, as shown in Fig. 37. The direct measurement
of force generation has been done at atmospheric pressure, which
diers greatly from downhole conditions. In order to predict the actual
force generation down-hole, back pressure has to be applied during
experiment. The axial oscillation generator tool discussed in (Ali et al.,
2011) provides a very high pressure drop (700 psi at 485 gpm and
10 lb/gal mud). The pressure drop for these tools ranges from 200 psi
to over 700 psi, depending on setup, and typically is in the 450600 psi
range (Gee et al., 2015). Gee et al. (2015) modeled the excitation as an
oscillating mass in the simulation, in the form of a sine wave. The Fig. 36. Frequency vs. ow rate of Agitator tool.
magnitude and frequency of excitations were 26 klbs (115.65 kN) and
20 Hz. Currently, an experimental facility is under development in the ADL at MUN that will be able to do this. From the experimental results,
the Agitator tool can be treated as a sinusoidal force source for
inclusion in the simulation model.

8. Simulation of horizontal drillstring with downhole tool

The main objective of this simulation is to show the ability of the


proposed model to capture the longitudinal and torsional motions of a
horizontal drillstring including predicting the eect of an axial excita-
tion tool on the motions of the drillstring. The model has the capability
to advance the bit and predict ROP. The simulation results for drilling
Fig. 32. Inlet pressure uctuation at 70 gpm ow rate. using a top drive and mud motor at 4340 m MD are shown in Figs. 38

281
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

Fig. 37. Pressure drop vs. ow rate of Agitator tool.

Fig. 39. The surface torque, bit speed and ROP for the case of without and with AES.

Fig. 38. The top drive speed, mud motor speed and WOB for the case of without and
with axial excitation source (AES).

46. The top of the drillstring was rotated at 5.2 rad/s (or 50 rpm) while
the mud motor was rotated at 13.7 rad/s (or 131 rpm). The string was Fig. 40. The axial excitation force, displacements at dierent locations for the case of
pushed down until it touched the rock and the WOB built up to 100 kN. without and with AES.

Then, rotary motion was applied to drillstring. At the beginning of the


simulation, the bit does not rotate due to the high cutting torque and the higher TOB, which is shown in Fig. 39, caused by higher WOB.
wellbore friction torque. The bit does not move forward, friction drag Again the drill bit reaches a constant speed of 18.9 rad/s when the top
throughout the string decreases and WOB increases as shown in drive provides the extra torque to overcome the cutting torque. The
Fig. 38. As soon as the drill bit rotates the bit starts to move forward increase in surface torque can be seen in Fig. 39. The 30.7% increment
(Fig. 39), friction drag is increased, and WOB is decreased to 100 kN in ROP due to higher WOB introduced by the AES is shown in Fig. 39.
(Fig. 38). The surface torque required to overcome the cutting torque at The axial displacements and forces in the drillstring segments gener-
the bit and frictional torque throughout the drilling is shown in Fig. 39. ated from the AES have been shown in Figs. 40 and 41, respectively.
The absence of uctuation in the surface torque indicates the constant The displacements at 350 m and 650 m behind bit show small
rotation at the bit (Fig. 39). The constant WOB and bit speed provide a oscillations, which is a good indication that the vibrations from the
constant ROP that can be veried from the ROP plot in Fig. 39. AES are not transferred to the drill bit where they could cause damage.
An axial excitation source (AES), which is a sinusoidal force, has There exist optimum values of amplitude and frequency for the AES
been placed at 650 m behind the bit according to the chart in Appx. B force to achieve higher WOB, higher ROP and less oscillation at the
and the comparison with drilling without AES is shown in Figs. 3840. BHA and bit.
The amplitude and frequency of the force have been chosen as 200 kN A comparison study for dierent amplitudes and frequencies of AES
and 125 rad/s (20 Hz) for the simulation results in Figs. 3841. The force on the drillstring motions has been conducted to show how the
WOB plot in Fig. 38 shows the 32.5% increment in WOB compared to model could be used to optimize the ROP. The displacements generated
drilling without AES. The bit speed comes down from 18.9 rad/s due to at the AES segment for dierent applied forces are shown in Fig. 42.

282
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

Fig. 44. The displacements of 350 m behind bit segment for dierent amplitudes and
frequencies of applied forces.

Fig. 41. Axial forces at dierent locations for the case of without and with AES.

Fig. 45. The WOB for dierent amplitudes and frequencies of applied forces.

Fig. 42. The AES segment displacements for dierent amplitudes and frequencies of
applied forces.

Fig. 46. The ROP for dierent amplitudes and frequencies of applied forces.

higher oscillation to the bit that can damage the bit. Thus a very high
amplitude and very low frequency AES force should be avoided in order
to protect the BHA and drill bit. A very low amplitude and a very high
Fig. 43. The plot of 950 m behind bit segment displacements for dierent amplitudes
frequency AES force is also not a good choice as it does not increase the
and frequencies of applied forces.
WOB and ROP. The horizontal drillstring model described herein is
eective at predicting the eect of downhole tool parameters on drilling
The higher amplitude force generated higher amplitude displacement performance and vibration throughout the string. Determining the
at the AES segment. Also, a higher amplitude is found when higher desired outputs of such tools through simulation is a cost-eective way
frequency AES force is applied. The simulation results in Figs. 43 and of generating essential tool design criteria.
44 show that the displacement oscillation throughout the drillstring is
very sensitive to the AES force frequency. Higher displacement 9. Conclusions
throughout the drillstring can be achieved by applying a low frequency
AES force that increase signicantly WOB (Fig. 45) and ROP (Fig. 46). Development of a bond graph model of a horizontal oilwell drill-
On the other hand, a very low frequency of AES force can provide a string, capturing longitudinal and torsional motions, using a lumped

283
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

segment approach has been presented. The dynamic model accounts dynamic stresses. While not the focus of the simulation exercise in this
for wellbore stick-slip friction. The model incorporates a modied bit- paper, stress analysis for fatigue design of components, or forensic
rock interaction model that allows the drill bit to move forward for analysis of failed components, are other potential uses of the model. An
prediction of the ROP. Implementation of the model in 20Sim experimental drilling facility, currently under development, will be
commercial software, which allows block diagrams to be superimposed used to parameterize the model for various rock and bottom-hole
on the bond graphs, greatly facilitated inclusion of the coupled long- pressure conditions, thereby increasing its predictive ability for design
itudinal and torsional degrees of freedom due to bit-rock interaction. and optimization of the drillstring, vibratory tool, shock absorbers, and
The simulation time is very fast compared to high order nite-and controllers. Finally, the proposed model is limited to the longitudinal
discrete-element models, making the model suitable as a tool for design and torsional motions of drillstring. In future, the model will be
and sensitivity analysis. The torque and drag obtained from the extended to include lateral motions of the drillstring.
proposed model is in qualitative agreement with eld data. Lab
experiments show that an Agitator tool can be represented as a Acknowledgements
sinusoidal force. The application of high amplitude and low frequency
of axial excitation force in the horizontal portion of the drillstring can This work was done at the Advanced Drilling Laboratory (ADL) at
provide better weight transfer to the bit and increase rate of penetra- Memorial University of Newfoundland (MUN) in St. John's, Canada.
tion. A trade-o between WOB transfer and vibration amplitude must Financial support was provided by Husky Energy, Suncor Energy,
be managed in order to protect the BHA and drill bit. The proposed Newfoundland and Labrador Research and Development Corporation,
model can be used as a tool for predrilling analysis. The ability to and Atlantic Canada Opportunities Agency under AIF contract number
predict segment forces throughout the string allows for prediction of 781-2636-192044.

Appendix A

An overview of bond graph formalism

Bond graph is an explicit graphical tool for capturing the energetic structure of a physical system and uniquely suited to the understanding of
physical system dynamics. Because of the ability to provide concise description of complex systems the bond graph formulation can be used in
hydraulics, mechatronics, thermodynamic and electric systems. The bond graph language expresses a general class of physical systems through
power (eort and ow) interactions and the factors of power have dierent interpretations in dierent physical domains.
Table A1 expresses the generalized power (eort and ow) variables and energy (momentum and displacement) variables in some physical
domains. The generalized inertias and capacitance in bond graph (Karnopp et al., 1999) store energy as a function of the system state variables, the
sources provide inputs from the environment, and the generalized resistors remove energy from the system. The state variables are generalized
momentum and dis-placement for inertias and capacitances, respectively. Where the time derivatives of generalized momentum p and displacement
q are generalized eort e and ow f. The power-conserving elements allow changes of state to take place. Such elements include power-continuous
generalized transformer (TF) and gyrator (GY) elements that algebraically relate elements of the eort and ow vectors into and out of the element.
In certain cases, such as large motion of rigid bodies in which coordinate transformations are functions of the geometric state, the constitutive laws
of these power-conserving elements can be state modulated. Dynamic force equilibrium and velocity summations in rigid body systems are
represented by power-conserving elements called 1 and 0 junctions, respectively.
Sources represent ports through which the system interacts with its environment. The power-conserving bond graph elements - TF, GY, 1
junctions, 0 junctions, and the bonds that connect them - are collectively referred to as junction structure. Table A2 denes the symbols and
constitutive laws of sources, storage and dissipative elements, and power-conserving elements in scalar form. Bond graphs may also be constructed
with the constitutive laws and junction structure in matrix-vector form, in which case the bond is indicated by a double line.
Power bonds contain a half-arrow that indicates the direction of algebraically positive power ow, and a causal stroke normal to the bond that
indicates whether the eort or ow variable is the input or output from the constitutive law of the connected elements. The constitutive laws in
Table A2 are consistent with the placement of the causal strokes. Full arrows are reserved for modulating signals that represent powerless
information ow such as orientation angles that determine the transformation matrix between a body-xed and inertia reference frame. A brief
introduction to the bond graph method is described in (Karnopp et al., 1999).

Table A1
Generalized bond graph quantities (adopted from Rideout et al., 2008).

Variable General Translation Rotation

Eort e (t ) Force Torque


Flow f (t ) Velocity Angular
Velocity
Momentum p = e dt Linear Angular
momentum momentum
Displacement q = f dt Displacement Angular
displacement

Energy p q Kinetic Kinetic


E ( p ) = f dp E (q ) = e dq
potential Potential

284
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

Table A2
Bond graph elements (adopted from Rideout et al., 2008).

Variable Symbol Constitutive law Causality


(Linear) Constraints

Sources

Flow Sf f = f (t ) Fixed flow out

Eort Sc e = e (t ) Fixed effort out

Energetic elements

Inertia I f=
1 df
e dt e = I dt Preferred
I I integral

Capacitor C e=
1
f dt f = C de Preferred
C dt integral
C

Resistor R e = Rf None
R f=
1
e
R

Port elements

Transformer 1 2 e2 = ne1 Effort in-effort


TF
f1 = nf2 out or
Flow in-ow
out

Modulated e2 = n ( ) e1
transformer f1 = n ( ) f2

Gyrator 1
TF
2 e2 = nf1 Flow in-effort
e1 = nf2 out or
Eort in-ow
out

Modulated e2 = n ( ) f1
Gyrator e1 = n ( ) f2

Constraint nodes
1 junction e2 = e1 e3 One flow input
f1 = f2 = f3

0 junction f2 = f1 f3 One effort


e1 = e2 = e3 input

Appendix B

See Table B1 and Figs. B1 and B2.


Well information.
Oil Company: Canadian Natural Resources Ltd.
Well: CNRL HZ Septimus C9-21-81-19.
LSD: 09-21-081-19W6M.
Rig: Precision Drilling Rig # 322.

285
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

Table B1
Drillstring configuration chart.

No. Congurations

I (MD = 200 m) Bit + Motor + HWDP(200 m)


II (MD = 620 m) Bit + Motor + HWDP(620 m)
III (MD = 720 m) Bit + Motor + HWDP(620 m) + DP (100 m)
IV (MD = Bit + Motor + HWDP(620 m) + DP (1100 m)
1720 m)
V (MD = 2180 m) Bit + Motor + HWDP(620 m) + DP (1560 m)
VI (MD = Bit + Motor + Collar (50) + DP (300 m) + Agitator +
2300 m) DP(1950 m)
VII (MD = Bit + Motor + Collar (50) + DP (300 m) + Agitator +
2556 m) DP(2113 m) + HWDP(93 m)
VIII (MD = Bit + Motor + Collar (50) + DP (300 m) + Agitator +
3062 m) DP(2113 m) + HWDP(561 m) + DP(38 m)
IX (MD = Bit + Motor + Collar (50) + DP (600 m) + Agitator +
4340 m) DP(1819 m) + HWDP(561 m) + DP(1310 m)

Well Name: CNRL Hz Septimus C9-21 -81 -19


-100 200 500 800 1100 1400 1700 2000 2300
-100
-400
Vertical Depth (m)

-700
-1000
-1300
-1600
-1900
-2200
-2500
Horizontal Distances (m)
Fig. B1. Sketch of well prole.

Fig. B2. Drillstring congurations in dierent depths.

Appendix C

Simulation data

See Table C1.

286
M. Sarker et al. Journal of Petroleum Science and Engineering 150 (2017) 272287

Table C1
Data used in horizontal oilwell drilling simulation.

Parameters Values Parameters Values

Drillstring data Rock damping 1.5105 N-s/m


Swivel and derrick mass 7031 kg Surface elevation amplitude s0 0.001
Kelly length 15 m Bit factor, b 1
Kelly outer diameter 0.379 m Cutting coefficient , C1, C2 1, 1.35108, 1.9104
Kelly inner diameter 0.0825 m Frictional coefficient 0 , , , , 0.06, 2, 1, 1, 0.01
DP outer diameter 0.101 m (4 in) Threshold force, Wfs 10000 N
DP inner diameter 0.0848 m (3.34 in) Hydraulic data
SUB outer diameter 0.136 m (5.354 in) Mud fluid density, m 1198 kg/m3
SUB inner diameter 0.057 m (2.244 in) Mud flow rate, Q Qm +Qa sin(qt )
collar outer diameter 0.125 m (4.921 in) Mean mud flow rate, Qm 0.022 m3/s
collar inner diameter 0.060 m (2.362 in) Mud flow pulsation amplitude, Qa 0.002 m3/s
motor HS outer diameter 0.121 m (4.763 in) Freq. of variation in mud flowrate, q 25.13 rad/s
motor HS inner diameter 0.0 m (0.0 in) Equivalent fluid viscosity for fluid resistance to rotation e 30103 Pa-s
Drillstring material Steel Weisbach friction factor outside drill pipe or collar, a 0.045
Wellbore diameter 0.18 m (7.086 in) Weisbach friction factor inside drill pipe or collar, p 0.035
Drill bit-rock data Motor data
Bit type PDC (Single cutter) V, f , P 2300 V, 377 rad/s, 4 pole
Drill bit diameter 0.159 m (6.259 in) Lls & Llr 0.0032 H, 0.0032 H
Drill bit mass 65 kg Lm 0.14329 H
Rock stiness 1.16109 N/m Rs & Rr 0.262, 0.187
Bit type PDC (Single cutter) Jm, Rm 11.06 kg m2, 0.05

References Lesage, M., Falconer, I.G., Wick, C.J., 1988. Evaluating Drilling Practice in Deviated
Wells with Torque and Weight Data. SPE Drilling Engineering.
Margolis, D., 2004. Fixed causality slip-stick friction models for use in simulation of non-
Aadnoy, B.S., 1998. Friction analysis for long reach wells. Presented at SPE/IADC linear systems. J. Syst. Control Eng., 216.
Drilling Conference, Dallas, Texas, USA. Rabbi, S.F., Sarker, M., Rideout, D.G., Butt, S.D., Rahman, M.A., 2015. Analysis of a
Aadnoy, B.S., Andersen, K., 2001. Design of oil wells using analytical friction models. J. hysteresis IPM motor drive for electric submersible pumps in harsh Atlantic oshore
Pet. Sci. Eng.. environments. ASME In: Proceedings of the 34th International Conference on
Aadnoy, B.S., Fazaelizadeh, M., Hareland, G., 2010. A 3-dimensional analytical model for Ocean, Oshore and Arctic Engineering, St. Johns, Newfoundland, Canada from
wellbore friction. J. Can. Pet. Technol.. May 31 - June 5.
Ali, A.A., Barton, S., Mohanna, A., 2011. Unique Axial Oscillation Tool Enhances Rideout, D.G., Stein, J.L., Louca, L.S., 2008. Systematic assessment of rigid internal
Performance of Directional Tools in Extended Reach Applications. Presented at the combustion engine dynamic coupling. American Society of mechanical engineers
Brasil Oshore Conference and Exhibition, Macae, Brazil, June 14-17. Transactions. J. Eng. Gas. Turbines Power 130, 2.
Gee, R., Hanley, R., Hussain, R., Canuel, L., Martinez, J., 2015. Axial oscillation tools vs. Sarker, M., Rideout, D.G., Butt, S.D., 2012. Dynamic model of an oilwell drillstring with
lateral vibration tools for friction reduction-whats the best way to shake the pipe? stick-slip and bit-bounce interaction. Presented at 10th International Conference on
Presented at the SPE/IADC Drilling Conference and Exhibition, London, UK, March Bond Graph Modeling and Simulation, Genoa, Italy, July 8-11.
17-19. Sheppard, J.H., Estes, B.L., Keller, S.R., 1987. Designing well paths to reduce drag and
Jardine, S., Malone, D., Sheppard, M., 1994. Putting a damper on drillings bad vibration. torque. SPE Drilling Engineering, 344345.
Oileld Review, January. Skyles, L.P., Amiraslani, Y.A., 2012. Converting static friction to kinetic friction to drill
Johancsik, C.A., Friesen, D.B., Dawson, R., 1984. Torque and drag in directional wells further and faster in directional holes. Presented at IADC/SPE Drilling Conference
prediction and measurement. J. Pet. Technol., 987992. and Exhibition, San Diego, California, USA, 6-8 March.
Karnopp, D., 1985. Computer simulation of stick-slip friction in mechanical dynamic Tveitdal, T., 2011. Torque and Drag Analysis of North Sea Wells Using New 3D Model
systems. ASME J. Dyn. Syst., Meas. Control V. 107 (1), 647668. (Master's Thesis). Faculty of Science and Technology, University of Stavanger,
Karnopp, D., Margolis, D.L., Rosenberg, R.C., 1999. System Dynamics; Modeling and Stavanger, Norway.
Simulation of Mechatronics Systems 3rd ed.. John Wiley and Sons, Inc, New York, Yigit, A.S., Christoforou, A.P., 2006. Stick-slip and bit-bounce interaction in Oilwell
USA. Drillstrings. American Society of mechanical engineers Transactions. J. Energy
Leine, R.I., 2000. Bifurcations in Discontinuous Mechanical Systems of Filippov-type Resour. Technol. 128 (4), 268274.
(Ph.D. thesis). Technical University of Eindhoven, Netherland.

287

Das könnte Ihnen auch gefallen