Sie sind auf Seite 1von 17

DC163492 DOI: 10.

2118/163492-PA Date: 15-March-14

Stage:

Page: 5

Total Pages: 17

Real-Time Evaluation of Hole-Cleaning


Conditions With a Transient
Cuttings-Transport Model
E. Cayeux, SPE, IRIS; T. Mesagan, SPE, S. Tanripada, SPE, and M. Zidan, SPE, Statoil; and
K.K. Fjelde, University of Stavanger

Summary
During a drilling operation, a real-time analysis of surface and
downhole measurements can give indications of poor hole cleaning. However, it is not always intuitive to understand how and
where the cuttings are settling in the borehole because the transportation of cuttings and the formation of cuttings beds are largely
influenced by the series of actions performed during the operation.
With a transient cuttings-transport model, it is possible to get a
continuously updated prognosis of the distribution of cuttings in
suspension and in beds along the annulus. This information can
be of prime importance for making decisions to deal with and prevent poor hole-cleaning conditions.
A transient cuttings-transport model has been obtained by
integrating closure laws for cuttings transport into a transient drilling model that accounts for both fluid transport and drillstring
mechanics.
This paper presents how this model was used to monitor two
different drilling operations in the North Sea: one using conventional drilling and one using managed-pressure drilling (MPD).
Some unknown parameters within the model (e.g., the size of the
cuttings particles) were calibrated to obtain a better match with
the top-side measurements (cuttings-flow rate, active pit reduction
as a result of cuttings removal). With the calibrated model, the
prediction of cuttings-bed locations was confirmed by actual drilling incidents such as packoffs and overpulls while tripping out of
hole.
On the basis of the calibrated transient cuttings-transport
model, it is thereby possible to evaluate the adjustments of the
drilling parameters that are necessary to stop and possibly remove
the cuttings beds, thus giving the drilling team the opportunity to
take remedial and preventive actions on the basis of quantitative
evaluations, rather than solely on the intuition and experience of
the decision makers.

Introduction
During drilling operations, ensuring proper hole-cleaning conditions is extremely important. Otherwise, serious drilling problems
can occur such as stuck-pipe incidents or packoff situations,
which can lead to the fracturing of the formation and resulting
mud losses. The end result of poor cuttings transport is an increase
in nonproductive time. To predict how cuttings are transported,
there has been performed a vast amount of experimental work and
different attempts on developing appropriate cuttings-transport
models. An overview of some of the work that has been performed is given by Pilehvari et al. (1999). It has turned out to be
quite complex to describe the cuttings-transport process because
transport is influenced by many different parameters such as wellbore geometry, inclination, fluid density, rheology, rate of penetration (ROP), drillstring rotation, flow patterns, flow rate, and
cuttings size. One could divide the modeling approach into two
C 2014 Society of Petroleum Engineers
Copyright V

This paper (SPE 163492) was accepted for presentation at the SPE/IADC Drilling
Conference and Exhibition, Amsterdam, 57 March 2013, and revised for publication.
Original manuscript received for review 25 March 2013. Revised manuscript received for
review 17 December 2013. Paper peer approved 10 January 2014.

main classes: the empirical approach in which correlations are


developed on the basis of experimental data (e.g., Larsen et al.
1997) or the use of more-fundamental mechanistic approaches
(Gavignet and Sobey 1989). The best approach still may not be
decided. The models developed are usually based on steady-state
considerations. However, in drilling operations, there are transient
phenomena taking place regularly that affect the cuttings transport
in the well as well as cuttings-bed evolution. Examples of such
phenomena are connections, temperature changes, variations in
ROP and flow rate, and drillstring-rotation adjustments. To describe the cuttings-transport process in an adequate manner, it
should be associated with the transient conditions taking place in
the well. In this paper, we have used an empirical correlation for
cuttings transport to build a transient drilling model that accounts
for both fluid transport and drillstring mechanics. The chosen cuttings-transport model is the one developed by Larsen et al. (1997)
that includes the extensions proposed by Jalukar (1993) and Bassal (1995). One can also note that the inclusion of such a model
will also improve the predictive capability of the transient drilling
model, because the physical processes taking place are interdependent. The combined model has been validated against two
field cases, and the results of this will be presented in the paper.
In comparing drilling models with real operational data, there is
often a requirement to perform some sort of calibration of the
models to improve the predictive capability. In this work, the sensitivity of the results to the cuttings-particle size and a bed-erosion
factor has been analyzed to determine in which conditions the output of the model matched the observations made during these drilling operations.
Cuttings-Transport Model
Transient Drilling Hydraulics. The transport of cuttings is governed by the mass-conservation equation, which in three dimensions can be expressed as
@q
r  q~
v 0; . . . . . . . . . . . . . . . . . . . . . . . .
@t

where t is time, q is the fluid density, and ~


v is the fluid-velocity
vector. This equation can be used to model the flow through the
entire wellbore, given appropriate boundary conditions. In this 3D
formulation, one can typically associate flow conditions at the
inlet of the pipes (typically the pump rate) and a zero flow at the
various walls (pipe wall or openhole/casing wall). Note that when
cuttings are produced by the drilling process, or when those cuttings exit the hydraulic system by settling into a bed, one has to
update the flow boundary conditions accordingly.
However, a 3D resolution of the problem is computationally
expensive and difficult to solve, and to analyze the transport of
cuttings inside the annulus at a macroscopic scale, it is sufficient
to restrict the analysis to a 1D problem. This can be performed by
cross-sectional averaging of the main quantities. When doing so,
the various flows through the pipe wall (e.g., cuttings beds, formation fluid gains, and losses) are no longer treated as boundary conditions, but as source terms in the conservation equation.
Typically, a drilling fluid is composed of a liquid phase (a solution such as a brine or an emulsion of water in oil) with several

March 2014 SPE Drilling & Completion


ID: jaganm Time: 16:19 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

Stage:

Page: 6

Total Pages: 17

Drilling fluid with


cuttings exiting the
control volume

Cross section
Wellbore

Control volume

Cuttings particle
setting in a bed

Drilling fluid with


cuttings entering the
control volume

Cuttings particle
returning to suspension
Drill-pipe
Cuttings bed

Fig. 1The mass exchange happens at the entry, exit, but also potentially along the control volume through cuttings beds.

types of solids in suspension: low-gravity solids (e.g., bentonite),


high-gravity solids (e.g., barite), and cuttings; the last component
is relevant only within the annulus. It should also be noted that
gas could be present in the drilling fluid, but for the sake of simplicity, the presented mathematical description will not include
any gaseous phase. One common formulation of the mass-balance
equation in the multiphase case is the drift-flux formulation that
considers that each phase is circulated in separate tubes with an
area that is given by the volume fraction of each phase (see Trapp
and Riemke 1986 or Liles and Reed 1978). By assuming common
pressure for each phase, one uses the mass-conservation equation
for each component:
8i 2 W;

X
@
@
fi Aqi fi Aqi vi
Cil ; . . . . . . . . . 2
@t
@s
l

where s is the curvilinear abscissa; A is the cross-sectional area of


a fluid element; the index i denotes the phase; the variables
fi ; qi ; vi are, respectively, the volume fraction, density, and velocity of phase i; the term Cil represents the interphase mass
exchange, with the convention that Cii m_ i , and m_ i is the mass
flux (per unit length) of phase i through the pipe walls.
This set of equations applies to all phases of the fluid. But looking at these equations from the perspective of cuttings transport,
one can see that the volume fractions, densities, velocities, interphase mass exchanges, and cross-sectional area vary with the position in the well and with time because cuttings are transported in
suspension, deposited in a cuttings bed, or returned to suspension.

MD (m)

t+7s

It should be noted that the presence of a cuttings bed changes


the available cross-sectional area for the passage of the fluid in
the annulus. Typically, it is considered that A changes with time
because of the movement of drillstring elements. For instance, at
any place along the drillstring where there is a change of the outside diameter in a drillstring element, the axial movement of the
drillstring causes a variation of the annulus cross-sectional area at
the position where this change of diameter in the drillstring element occurs. In addition, it should be noted that the variation of
cuttings-bed size through time also affects the cross-sectional area
(Fig. 1).
Furthermore, to calculate the distribution of pressure and temperature in the system, we need to consider the exchange of momentum and heat (Lage et al. 2003). The momentum balance,
derived from the Navier-Stokes equation, is written as

@
@ 
At; sqm t; svm t; s
At; sqm t; sv2m t; s
@t
@s
@
At; s pt; s At; sfKt; s  qm t; sgcos#sg;
@s
                   3
where p is the pressure, K is the frictional pressure-loss term, # is
the average inclination of the control-fluid element, and g is the
gravitational acceleration of the Earth. Eqs. 2 and 3 describe the
fully transient behavior covering both pressure-pulse propagation
and mass transport, as illustrated in Fig. 2. In this figure, the mud

t+14s

t+21s

t+28s

500
1000
1500
2000
2500
3000
3500
4000
4500
5000
5500
0 5 10 15 20 1 1.2 1.4 1.6 1.8 0 5 10 15 20 1 1.2 1.4 1.6 1.8
Fluid velocity
Fluid velocity
ECD (sg)
ECD (sg)
(m/s)
(m/s)

0 5 10 15 20 1 1.2 1.4 1.6 1.8


Fluid velocity
ECD (sg)
(m/s)

0 5 10 15 20 1 1.2 1.4 1.6 1.8


Fluid velocity
ECD (sg)
(m/s)
In annulus
In drill-string

Fig. 2Transient hydraulic effects during the acceleration of the mud pumps.
6

March 2014 SPE Drilling & Completion


ID: jaganm Time: 16:19 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

Casing #1
Fluid in annulus with cuttings
Fluid in drill-string
Drill-string
Cuttings bed

Formation
Cement #2
Casing #2
Cement #1

Control Volume
Fig. 3The different layers participating in the heat exchange
in a control volume.

pumps are accelerated smoothly from 0 to 1500 L/min. One can


see how the fluid-velocity wave propagates first in the drillstring
and then in the annulus. We can also see how the downhole pressure gradient increases as a consequence of the fluid movement
when the velocity wave propagates in the annulus.
Finally, the energy conservation as a result of heat transfer can
be written as (Marshall and Bentsen 1982):
@
q t; sHt; s  rQf t; s Qc t; s  qs t; s 0;
@t m
                   4
where H is the enthalpy per mass unit, Qf is the forced convective
term, Qc is the conductive and natural-convective term, and qs is
the heat generated by mechanical and hydraulic frictions.
The forced convective term can be expressed as
Qf qm t; sHt; svm t; s: . . . . . . . . . . . . . . . . . . . . 5
The conductive and natural-convective term does not have a
general expression. In the case of purely convective isotropic material, we can use
Qc kt; srTt; s; . . . . . . . . . . . . . . . . . . . . . . . . . 6
where k is the thermal conductivity and T is the temperature. The
heat exchange in a control volume is occurring between different
layers: fluid inside the drillstring, drillstring walls, fluid in the
annulus, casing walls, cement, and formation (Fig. 3). It is therefore a 2D model in a curvilinear/cylindrical coordinate system
with an assumption of uniformity for the angular dimension. Note

Stage:

Page: 7

Total Pages: 17

that the thermal conductivity of the mud is affected by the concentration of cuttings. Furthermore, the presence of cuttings beds
is also changing the thermal conductivity in the annulus because
heat transfer can happen between the wall of the drillstring and
the casing or formation through the cuttings bed instead of the
annulus fluid.
A partial-differential equation (Eq. 4) describes the transient
variations of the drilling-fluid temperature in the drillstring and in
the annulus, as illustrated by Fig. 4. Because of the slow evolution
of heat-transfer processes and the relatively rapid change in drilling
parameters, steady-state conditions are, in practice, never reached.
Cuttings in Suspension
The presence of cuttings in suspension in the drilling fluid
changes the properties of the original drilling fluid. This affects
the density, viscosity, specific heat capacity, and thermal conductivity of the mud during the transport of the cuttings.
Drilling-Fluid Density. Because of its compressibility and thermal expansion, the density of each of the components of the drilling fluid changes as a function of pressure and temperature.
Because drilling fluids are used in a wide range of pressures and
temperatures, it is seldom acceptable to consider the compressibility and thermal expansion of the fluid components as constant.
For instance, Isambourg et al. (1996) recommended a biquadratic
relationship for the density of the components with respect to temperature and pressure. In this model, the liquid phase will be considered compressible whereas the solid phase will be treated as
incompressible. The liquid phase can be either a solution (brine or
diluted cesium formate) or an emulsion (a colloid system such as
water dispersed in oil). The density variations of a brine as a function of temperature and pressure depend on the type of salts and
their concentration in the electrolyte (Fig. 5). For instance, Kemp
et al. (1989) gave a precise description on how to calculate the
density of a brine containing different salts at various concentrations. Similarly, the density of base oils used in invert emulsions
depends on temperature and pressure (Fig. 6). Finally, it should
be noted that the solid phase can have, at most, three components:
high-gravity solids, low-gravity solids, and cuttings.
The density of the mixture is given by the combination of the
densities of the components weighted by the volume fraction of
each element:
X
X
fi qi p; T; with
fi 1; . . . . . . . . . 7
qm p; T
i2X

i2X

where qm is the density of the mud, p is the pressure, T is the temperature, X is the set of indices for the different components (g

0
30 in. at 423 mMD
26 in. at 561 mMD

MD (m)

500
1000

20 in. at 1420 mMD

1500

13 3 8 in. at 1783 mMD


2000
10 in. at 2239 mMD
2500
0

20 40 0
20 40 0
20 40 0
20 40 0
20 40 0
20 40
Temperature Temperature Temperature Temperature Temperature Temperature
(C)
(C)
(C)
(C)
(C)
(C)
t

t+20 min

t+40 min

t+60 min

t+80 min

t+100 min

9 5 8 in. at 2247 mMD


8 in. openhole

Temperature in annulus
Temperature in drill-string
Geo-thermal temperature

Fig. 4Evolution of the temperature when establishing circulation from geothermal conditions.
March 2014 SPE Drilling & Completion
ID: jaganm Time: 16:19 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

1,14

NaCl 20C 10% wt

1,12

Page: 8

Total Pages: 17

Base Oil Density

0,85

CaCl2 20C 10% wt

0,8

1,1

NaCl 50C 10% wt


CaCl2 50C 10% wt

1,08

KCl 50C 10% wt

1,06

NaCl 80C 10% wt

1,04

Density (s.g.)

KCL 20C 10% wt

Density (s.g.)

Stage:

KCl 80C 10% wt

200

400
600
800
Pressure (bar)

93,3C

0,7

176,7C

0,65

CaCl2 80C 10% wt

1,02

23,9C

0,75

0,6
0

1000

200

800
400
600
Pressure (bara)

1000

1200

Fig. 5Brine-density variations for different dissolved salts for


a concentration of 10% in weight.

Fig. 6Example of the density variations of a typical base oil


as a function of pressure and temperature.

for gas, w for water or brine, o for oil, lgs for low-gravity solid,
hgs for high-gravity solid, and c for cuttings), fi is the volume
fraction of the ith component, and qi is the density of the ith
component.
Because of the mutual compressibility of the liquid phases, the
volume fractions of each phase depend on the initial concentration
and the current conditions of temperature and pressure at the
depth of investigation [see Cayeux and Lande (2013) for a
detailed description of the calculation of the volume fractions].
To exemplify this aspect of the variability of volume fractions at
conditions of pressure and temperature different from the initial
ones, we will look in more detail at the cuttings volume fraction.
If we consider that the volume fraction of cuttings fc1 is known at
a given temperature T1 and pressure p1 , then to determine the volume fraction of solids at a different pressure and temperature, we
should consider that the solid phase is actually incompressible but
the dispersion medium (oil/brine) is actually compressible and
dilatable. With changes of pressure and temperature, the volume
occupied by the dispersion medium reduces or expands. As a consequence, the distance between the solid particles changes and,
therefore, so does the volume fraction of the solids (Fig. 7).
The volume fraction of solids can be expressed as a function
of the initial volume fraction and the local temperature and pressure as follows:

Rheology. Drilling muds are non-Newtonian fluids; they are


more precisely shear-thinning fluids with a yield stress. A typical
rheological model that describes such liquids satisfactorily is that
of Herschel and Bulkley (s s0 K c_ n , with s0 the yield stress, K
the consistency index, and n the flow-behavior index), but it
yields mathematical formulations for the fluid flow in pipes and
annuli that are not readily solvable. However, a rheological formulation has been proposed (Robertson and Stiff 1976) that fits
well with drilling-fluid rheometer measurements (see the measurements made by Beirute and Flumerfelt 1977). This allows for
analytical expressions of the shear rate, and solvable mathematical formulations of the pressure loss in laminar flow inside a tube
or a concentric annulus (Whittaker 1985), which, in turn, help
improve the computational efficiency of algorithms that are used
for transient-flow modeling.
The Robertson-Stiff rheology is defined by

fc p; T

fc1 qm p; T
; . . . . . . . . . . . . . 8
1  fc1 qm1 fc1 qm p; T

where qm1 is the density of the mud at the reference pressure p1


and temperature T1 . Fig. 8 shows how the local density of the
mud in the annulus is affected by the presence of cuttings. In addition, the local pressure and temperature at a given depth also
influence the density of the mud. It should be noticed that when
there is circulation and the bit is off-bottom, the cuttings already
under transport are still moving up while no new cuttings are produced at the bottom hole, thus causing large variations in the local
mud density as a function of depth.

s A_c CB ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
where s is the shear stress, c_ is the shear rate, and A, B, and C are
the coefficients of the model. This is the model that is used in the
described transient hydraulic model. It should be noted that, as with
any other fluid, the rheology of drilling mud depends on temperature. In addition, the mud viscosity increases exponentially with
larger pressures, which is true at any temperature (Houwen and
Geehan 1986), following an Arrhenius-type law (Figs. 9 and 10).
Furthermore, the cuttings in suspension modify the effective
viscosity of the mud. The effect of solid particles in suspension on
the rheology of fluids was first studied by Einstein (1906). This
analysis is based on an energy balance to determine the viscosity
of a suspension of solids in Newtonian liquids:
gmc p; T gm p; T1 2:5fc p; T; 8fc 2 0; 0:54;
                   10
where gmc is the viscosity of the drilling fluid with the cuttings, gm
is the original viscosity of the mud, and fc is the volume fraction

Volume of dispersion
Medium reduces

Control volume
Concentration of dispersed
Phase increases
Pressure
Fig. 7Effect of compression on the solid volume fraction.
8

March 2014 SPE Drilling & Completion


ID: jaganm Time: 16:20 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

MD (m)

Stage:

Page: 9

Total Pages: 17

t+20 min

t+10 min

t+30 min

500
1000
1500
2000
2500
3000
3500
4000
4500
5000
5500
0

1.17 1.19

2
4
1.17 1.19
0
2
4
Cuttings
Density (sg)
Cuttings
Concentration
Concentration
(%)
(%)

Density (sg)

Cuttings
Concentration
(%)

1.17 1.19
Density (sg)

2
4
Cuttings
Concentration
(%)

1.17 1.19
Density (sg)

In annulus
In drill-string
Fig. 8The local density of the mud increases with the concentration of cuttings. During a period of circulation without production
of cuttings, the local mud density is back to normal.
25C, 200 bar
25C, 100 bar and 150 bar
25C, 50 bar
25C, 1 bar

25

50C, 200 bar


50C, 150 bar
50C, 50 bar and 100 bar
50C, 1 bar
75C, 200 bar
75C, 100 bar and 150 bar
75C, 1 bar

Shear stress (Pa)

20

15

10

200

400
600
Shear rate (s1)

800

1,000

Fig. 9Temperature and pressure dependence of the rheology of a water-based mud measured with an Anton Paar scientific
rheometer.
Temperature and pressure dependence of flow curves of an OBM with density 1.6 sg at 50C
60
25C, 200 bar
25C, 150 bar
25C, 100 bar
25C, 1 bar
50C, 200 bar
50C, 150 bar
50C, 100 bar

Shear stress (Pa)

50

40

50C, 1 bar

30

20

10

0
0

200

400

600

800

1,000

Shear rate (s1)

Fig. 10Temperature and pressure dependence of the rheology of an oil-based mud measured with an Anton Paar scientific
rheometer.
March 2014 SPE Drilling & Completion
ID: jaganm Time: 16:20 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

of cuttings in the mud. Hastchek (1910) proposed a modification


of Einsteins equation on the basis of Stokes law for a slowly
falling ball in an infinite continuum. One major advantage of this
model is that it is valid for any concentrations of particles:
gmc p; T gm p; T1 4:5fc p; T; 8fc 2 0; 0:74
1
gmc p; T gm p; T
; 8fc 2 0:74; 1:
1
1  fc =3 p; T
                   11
The viscosity of the original mud can be derived from the
rheological constitutive law of the drilling fluid by use of
s
gm ; 8_c 6 0: . . . . . . . . . . . . . . . . . . . . . . . . .
c_

12

Stage:

Cp m

X
i2X

ei Cpi ; with 8i 2 X; ei

fi qi
; . . . . . . . . . . . 13
qm

where Cpm is the specific heat of the drilling mud; ei ; i 2 X are the
mass fractions of the components; Cpi ; i 2 X are the specific heat
capacities of each components; and X is a set of indices representing each component (w for water or brine, o for oil, lgs for lowgravity solid, hgs for high-gravity solid, and c for cuttings). The
mass fractions are derived from the volume fractions by use of the
density of the components.
However, the specific heat capacity of the different components
is seldom well-known. This leads to a substantial uncertainty on the
estimated specific heat capacity of the mix. Ideally, measurements
taken either at the laboratory or at the rigsite should be performed
at regular intervals to verify the accuracy of the estimations.
Thermal Conductivity. The calculation of the effective thermal
conductivity in a heterogeneous multicomponent medium is not
simple. The first solution to the calculation of the effective thermal conductivity of a suspension of solid particles in a homogeneous medium was described by Maxwell (1873) and was based on
the assumption that the particles can be assimilated to spheres and
that their concentration is small. For normal weighted drilling fluids, the concentration of solid particles cannot be considered
small. Rayleigh extended Maxwells model to a higher order of
concentration by considering spherical particles, yet nontouching,
but distributed on a regular cubic lattice. Rayleighs model, less
and less accurate when the mass fraction of particles in suspension reaches 0.5236, was modified by Churchill to circumvent
that problem (Kandula 2011). The resulting expression of the
effective thermal conductivity for the mud is then
2k
6 3k 73
3  3k 103
 2fs 0:409
fs  2:133
fs
1

k
4

3k
4 3k ;
km k l
2k
6 3k 73
3  3k 103
fs 0:409
fs  0:906
fs
1k
4 3k
4 3k
                   14
ks
is the solid-to-liquid thermal-conductivity ratio.
kl
The background continuous medium (i.e., the liquid phase)
can be a mixture solution (typically oil and brine) of different fluids with different thermal conductivity. A first approximation of
the effective thermal conductivity of the liquid mixture is
X
1i ki ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
kl
where k

i2H

where H is a set of indices for the liquid components (typically o


for oil and w for water/brine), fi is the mass fraction of the ith
component relative to the mass of the liquid mix, and ki is the
10

Total Pages: 17

thermal conductivity the of the ith component of the liquid. Rowley (1981) reported that experimental measurements showed that
the weighted average method resulted in an excess thermal conductivity. This can be compensated for by adding a correction
term in the weighted average formula:
"
#
X
X X 1j Gij kji  ki
1 i ki
1i
; . . . . . . . 16
kl T
Ri2H 1i Gij
i2H
i2H
j2H
where
aAij

8i; j 2 H; Gij e RT and kji

1i 1ii ki 1j 1jj kj


. . . . . . 17
1i 1ii 1j 1jj

in which
8i 2 H; 1i

Specific Heat Capacity. The theoretical specific heat capacity of


the multicomponent medium is the weighted average in terms of
mass fractions of the specific heat capacity of each component:

Page: 10

1j Gji
1i
and 8i; j 2 H; 1ji
;
Rj2H 1j Gji
Rk2H 1k Gki
                   18

where a is the nonrandomness factor of the nonrandom two-liquid


(NRTL) model, Aij are the NRTL parameters, and R is the universal gas constant. These parameters can be found in the literature
for mixes of two liquids (Nagata 1973; Gmehling and Onken
1977) or from the DDBST (2014).
To calculate the effective thermal conductivity of the mud
with several types of particles, one should, rigorously, use an
extension to the Churchill model to n dispersed phases embedded
in one continuous medium. Such an extension exists for the Maxwell model (made by Eucken in 1940), but it does not exist for the
Churchill model. Even though it is not exactly correct, it is possible to use, as a practical approximation, the Churchill formula,
recursively taking one component at a time and using the resulting
thermal conductivity as the thermal conductivity of the background medium when adding the next component.
As with the estimation of the specific heat capacity, this
approach depends on knowledge of the thermal conductivity of
each component, which is seldom the case. The measurement of
field samples with accurate instruments, at the wellsite or in the
laboratory, could alleviate this uncertainty.
Cuttings Transport and Settling
A cuttings particle in suspension within a moving drilling fluid is
subject to several forces. The first set of forces is the static ones:
the gravity force (Fg) and the buoyancy force (Fb). In addition,
there is a dynamic force: the frictional force (Clark and Bickham
1994). The frictional force is decomposed into a drag force (FD)
following the flow direction and a lift force (FL) perpendicular to
the flow movement (Fig. 11). The drag and lift forces depend on
the local velocity of the fluid around the cuttings particle.
If the fluid velocity is large enough, the motion of a cuttings
particle can be mostly in the axial direction of the borehole. To
obtain such a direct transport, the lift force has to balance the components of the gravitational and buoyancy forces in the perpendicular plane to the wellbore axis. In that case, we do need only to
consider a bulk fluid velocity (averaged over the whole cross section), and at a constant fluid velocity, the cuttings particle reaches a
terminal velocity (the slip velocity) compared with the fluid. If the
slip velocity is positive, then the cuttings particle is transported.
Rayleigh derived an expression of the drag force that is quadratic in velocity:
1
8NRe > 1; 000; FD qm v2s Cd A; . . . . . . . . . . . . . . . .19
2
where Cd is the drag coefficient, vs is the slip velocity compared
with the fluid flow, and A is the reference area (i.e., the orthographic projection of the object on a plane perpendicular to the
direction of motion). The drag coefficient of this formula has a
constant value only for objects that have a blunt form factor and
when the particle Reynolds number (NRe ) is large enough to produce turbulence behind the object (NRe > 1; 000). The drag coefficient of a rough sphere at high particle Reynolds number is
March 2014 SPE Drilling & Completion

ID: jaganm Time: 16:20 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

Stage:

Longitudinal view

ile

ud

l
ve

FD

Fb

FL
Cuttings particle

Mud velocity profile

y
cit

Total Pages: 17

Cross-section

of

pr

Page: 11

Cuttings particle

FL

Fb

Fg

ril

l-p

ut

ip

tin

gs

Fg

vm

be

Rotating
drill-pipe

vm

Tangential
component
of mud flow

d
Cuttings bed

Fig. 11Forces acting on a cuttings particle in suspension.

approximately 0.40 to 0.45 (Loth 2008). Note that the particle


Reynolds number is defined by
NRe

dvs qm
; . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
gm

where d is the diameter of the particle and gm is the effective viscosity of the fluid.
At lower particle Reynolds number, the drag coefficient is no
longer constant. It is possible to find drag-coefficient correlations
for irregularly shaped particles (Cdshape ) in a Newtonian fluid (Holtzer and Sommerfeld 2008), for spherical particles moving in a
shear-thinning (Cdthin ) fluid (Renaud et al. 2004), and for spherical
particles displaced by a viscoplastic (Cdplast ) fluid (Beaulne and
Mitsoulis 1997). However, there is not yet any published correlation for a drag coefficient that accounts simultaneously for the
particle shape and the non-Newtonian rheological characteristics
of the fluid. To generalize the applicability of the preceding correlations, it is possible to use a superposition principle of the effects
caused by the particle shape, shear thinning, and viscoplasticity.
With such a superposition principle, an expression for the effective drag coefficient (Cdeff ) can be written as
Cdeff Cdshape Cdthin n  Cdthin 1 Cdplast Bi  Cdplast 0;
                   21
where n is the flow-behavior index of a Herschel-Bulkley rheology and Bi s0 n is the Bingham number.
K

vs
dp

With the estimation of the effective drag coefficient, we can


derive the terminal velocity by solving the force equilibrium:
p
p
p
 d3 qp gcos# d3 qf gcos# d 2 qf v2s Cdeff
6
6
8


4
qc
0 () dgcos#
 1 v2s Cdeff ; . . . . . . . . . 22
3
qm
where qc is the density of the cuttings particle, g is the gravitational acceleration of the Earth, and # is the inclination. By solving this equation in vs , it is possible to estimate the slip velocity,
in the axial direction, of an irregular particle moving into a shearthinning fluid with yield stress.
When the fluid flow is insufficient to transport the cuttings particles mostly in the axial direction of the borehole, it is no longer
possible to use the assumption of a bulk-fluid velocity. In this
more general case, it should be noted that the fluid-velocity field
is not uniform in the cross section of the annulus, and close to the
walls or on the low side when the drillpipes are decentered (i.e.,
for an inclined section), the local fluid velocity may therefore not

be able to transport the cuttings even though it is sufficient on the


high side of the annulus. As a consequence, the cuttings particles
that do not enter the main fluid stream are not transported, and
they then settle into a bed. If the drilling-fluid flow is turbulent, it
is likely that such cuttings-bed particles may be lifted in the main
fluid stream and therefore be carried away. A similar effect can be
obtained by the stirring effect caused by the rotation of the drillstring. Such a rotation causes a secondary velocity field for the
drilling fluid, perpendicular to the axis of the annulus, which
should be combined to the axial velocity field. This can lift cuttings particles from the low side into the main fluid stream. The
rotation of the drillpipe may cause local turbulences that do not
necessarily exist in the main stream. This turbulence effect also
can help the particles move into the main fluid course where they
can then be sent downstream. So, a method based on terminal particle velocity is valid only when the fluid velocity across the section is relatively symmetric around the wellbore axis (i.e., little
effect of drillpipe eccentricity and drillpipe rotation). These conditions correspond to the vertical or low inclined part of the well.
Experimental studies indicate that the inclination limit is approximately 358 (Tomren et al. 1986).
Experiments have confirmed that the eccentricity of the drillpipe inside the annulus and the drillstring rotational velocity play
an important role in the actual cuttings transport for the moreinclined part of the hole. Nazari et al. (2010) give an exhaustive
overview of the dominant factors influencing the transport of cuttings, including the two mechanisms mentioned previously.
Unfortunately, according to Ozbayoglu et al. (2008), a mechanistic relationship, including both rotation and inclination, has not
yet been published.
There are three possibilities to account for this complex phenomenon: Use an empirical model based on relevant experimental
data, develop a physical law that uses statistical physics methods,
or synthetize the results of many simulations made with a detailed
computational-fluid-dynamics (CFD) model of the problem, similar to the work performed on annular pressure loss around stabilizers presented by Yao and Robello (2008). In a first approach, we
have chosen a method that is based on empirical correlations
because of readily available results.
Indeed, starting in the early 1980s, several experiments have
been performed to develop correlation models of the minimal critical fluid velocity necessary to transport cuttings in various conditions. Ford and Peden of the Heriot-Watt University, were among
the first to present their results (Ford et al. 1990; Peden et al. 1990).
Shortly after, the University of Tulsa started an extensive and systematic series of experiments in a large-scale flow loop. From the
experimental data, they extracted an empirical model first described
by Larsen et al. (1997). This initial model was modified a first time
to better account for the effect of hole size (Jalukar 1993), and then

March 2014 SPE Drilling & Completion


ID: jaganm Time: 16:20 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

11

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

a second modification was made to include the effect of the drillpipe rotational velocity, because this effect was not accounted for
in the initial study (Bassal 1995). A few years later, another independent study was presented by the Bandung Institute of Technology, Indonesia (Rubiandini 1999).
A systematic analysis of the models proposed by Larsen and
Rubiandini has shown that Rubiandinis model, in all tested configurations, provides larger critical transport fluid velocities
(CTFVs) than that of Larsen (Ranjbar 2010). In the work of Ranjbar, the mechanistic approaches of Gavignet and Sobey (1989)
and Kamp and Rivero (1999) were also described. As reported by
Kamp and Rivero (1999), their model gave quite small/unreasonable CTFVs. The Larsen and Jalukar model predicted cuttingsbed buildup at a flow rate that was ten times higher.
Practical use of the Larsens model in real drilling operations
has shown that this model is usually conservative, but the two
extensions made by Bassal and Jalukar have not been accounted
for, either in the referred practical drilling operations or in the
analysis made by Ranjbar. However, those extensions are quite
central to model the effect of drillpipe rotation in various hole
sizes and should not be disregarded.
In this paper, we have used Larsens model, combined with
the extensions of Jalukar and Bassal, to calculate the CTFV along
the annulus. The resulting correlation model accounts for the
effects of inclination, fluid velocity, fluid rheology, mud weight,
cuttings size, cuttings concentration, drillpipe eccentricity, drillpipe rotational velocity, and ratio between hole size and drillpipe
size. It should be noted that this model is based on approximately
2,000 experiments. However, the inclination used while acquiring
the data has been varied between 55 and 908. Furthermore, the
correlation on the drilling-fluid apparent viscosity is expressed
only through a Bingham plastic rheology [all the experiments
have been made with water-based muds (WBMs)]. With a similar
rheology, the type of mud has little impact on the CTFV (Hareland et al. 1993). However, cuttings-bed erosion behaves differently in oil-based mud (OBM), compared with WBMs. Because
the extended Larsens model is used only for calculating the
CTFV, the lack of correlation with mud type should not affect the
results too much. Even though the extended Larsens model has
been developed with data from experiments with an inclination
higher than 558, the estimation of the CTFV is considered to be
valid from an inclination angle as low as 358.
It should be noted that the CTFV value calculated by the
extended Larsens model corresponds to conditions in which there
is no cuttings bed forming at all. Therefore, when the local fluid velocity is lower than the CTFV, it can be expected that cuttings will
settle. In such conditions, for a given control volume, the concentration of cuttings in suspension shall be updated, as well as the
height of the bed. In the opposite case, if the fluid velocity is greater
than the CTFV, the cuttings that were trapped in a cuttings bed
return to suspension. This also has a consequence on the local (i.e.,
at each depth along the annulus) concentration of cuttings in suspension and the cuttings-bed height. In cases in which several particle sizes are used, the depositing and resuspension of cuttings
particles may concern some of the cuttings particles but not all
them. It is, therefore, necessary to account for the local cuttingssize distribution of the particles in suspension and those lying in a
bed. Cuttings particles continue to accumulate in a bed until the
free cross-sectional area has been reduced in such a way that the
fluid velocity is larger than the CTFV. At that moment, the cuttings-bed height limit has been reached (Clark and Bickham 1994).
To calculate the size of the cuttings bed in the annulus at a
given depth, we need to account for the actual packing efficiency
of particles in the bed (i.e., the ratio of the actual particle volume
to the occupied volume). For the monodispersed packing problem
p
(a single size), the maximal packing efficiency is p  0:74048,
18
which is the highest possible density among all possible lattice
packing, as demonstrated by Gauss (1831). In practice, when the
spheres are added randomly, the packing is irregular, and the
maximal achievable density is lower than the best lattice packing.
It has been demonstrated that, with jammed packing, the packing
12

Stage:

Page: 12

Total Pages: 17

efficiency cannot exceed the limit of 63.4% in the most-compact


way and 55% for loose packing (Song et al. 2008). The polydispersed (n-components mixture) packing problem is extremely
complex. In that case, we consider that the different particle sizes
are stacked on top of each other, with the largest particle at the
bottom and the smallest at the top. However, for the binary hardsphere packing problem (two sizes) with coarse and fine particles,
there exist solutions for calculating the packing efficiency of the
mixture of particles. Zheng et al. (1995) have proposed

 5 4
PEmix PEc 1  PEc PEf eXf lnXf 4PEc e r ; . . . 23
where PEmix is the packing efficiency of the mix of coarse and
fine particles, PEc is the packing efficiency of the coarse particles,
PEf is the packing efficiency of the fine particles, Xf is the volume
fraction of fine particles, r is the size ratio of coarse particles to
fine particles, and e is Eulers number.
All these packing efficiencies are for a very large region so
that the effect of boundaries is insignificant. This hypothesis is
true as long as the cuttings size is small compared with the radius
of the borehole. Thereafter, it is possible to determine the area
occupied by cuttings (Ac ) at a given depth by accounting for the
packing efficiency. For a monodispersed system or a binary dispersion with coarse and fine particles,
Ac

npd3
; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
6LPE

where n is the number of particles in the control volume, d is the


diameter of the particles, PE is the packing efficiency, and L is
the length of the controlled volume. For the case of a multidispersed cuttings bed, the area occupied by the cuttings is
Xk ni pd3
i

Ac

i1

6PEi

; . . . . . . . . . . . . . . . . . . . . . . . . . 25

where k is the number of different particle sizes, ni is the number


of particles of the ith size in the controlled volume, di is the diameter of the particles of the ith size, and PEi is the packing efficiency for the particle of the ith size. Then, it is possible to find
the free area in the annulus:
Af prw2  Ac  prp2 ; . . . . . . . . . . . . . . . . . . . . . . . 26
where Af is the free area in the annulus cross section, rw is the radius of the wellbore, and rp is the radius of the drillpipe.
Finally, to find the height of the cuttings bed (Fig. 12), one
needs to solve the following piecewise equation:


rw  hc 2
rw
if hc  rw  e  rp ; Ac acos
rw
q
rw  hc rw2  rw  hc 2
if hc > rw  e  rp and hc  rw  e rp ; Ac


q
rw  hc 2
rw  rw  hc rw2  rw  hc 2
acos
rw



q

rw  hc  e 2
rp  rw  hc  e rp2 rw  hc  e2
 acos
rp


rw  hc 2
rw
if hc > rw  e rp ; Ac acos
rw
q
rw  hc rw2  rw  hc 2  prp2 :
                   27
During transportation, the cuttings size changes as a result of
mechanical interactions caused by grinding. This phenomenon is
clearly observable when using a dual concentric drillstring drilling
method (Belarde and Vestavik 2011). With this new drilling
method, a dual concentric drillpipe is used where cuttings transport takes place inside the inner drillpipe rather than the annulus.
March 2014 SPE Drilling & Completion

ID: jaganm Time: 16:20 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

rw

Stage:

Page: 13

Total Pages: 17

rw

rp

rw

rp

rp
hc

hc
hc
Fig. 12The different configurations for the cuttings bed and the drillpipe.

Alternating from conventional drilling to the dual-concentric drilling method within the same drilling operation, significant variations in size and shape of the cuttings (Fig. 13) were observed.
This process is poorly understood, and we have not attempted to
incorporate any modeling of cuttings-size variation in the presented model, even though it probably has important consequences on the cuttings transport and on the evolution of cuttings beds.
Cuttings-Bed Erosion
When a cuttings bed is formed, it is subjected to erosional forces
as a result of the flow of drilling fluid in the annulus. Laboratory
experiments have shown three possibly intertwined modes of
streambed erosion (Ramadan et al. 2001): individual-particle
entrainment at the surface of the bed (also called surface erosion),
bed-surface-layer displacement (called mass erosion), and entirebed fluidization, resulting in a bulk movement of the bed. In the
first mode of erosion, the movement of individual particles at the
surface of the bed is the result of forces acting on each particle. In
that case, there are basically three possibilities: The lift force, as a
result of mud flow, on the particle is large enough to temporarily
extract the particle from the bed; the rotation of the drillpipe creates an additional lift force that can start the temporary entrainment of the particle downstream; and the drag force, induced by
the fluid flow, enables the particle to roll on top of the neighboring
particles. Mass erosion occurs when the entire top layer of the cuttings bed starts moving because the shear stress across the bed is
larger than the cohesive strength of the structure, thus causing a
plane of failure separating the top part of the bed from the rest. In
the third mode of erosion, the whole bed behaves like a fluid.
We will focus on the surface-erosion mode. The mud flow
induced by pumping or by the axial movement of the drillpipe creates a fluid-velocity field in the axial direction of the wellbore that
is referred to as !
va . The rotation of the drillpipe creates a fluid-ve-

Fig. 13Comparison of cuttings sizes from conventional drilling and dual-concentric drilling method (courtesy of Belarde
and Vestavik 2011).

locity field perpendicular to the wellbore direction, at least if circular Couette flow can be assumed. If the Reynolds number is high
enough, then Taylor vortices will exist, resulting in much-morecomplex local velocities. For the moment, we will consider a laminar circular Couette flow and denote the tangential component of
the fluid velocity !
vt . The local fluid velocity v!
m is, therefore, the
sum of the tangential fluid velocity and the axial velocity
! !
(v!
m va vt ). The angle of the wellbore axis is denoted # and is,
in fact, the inclination of the borehole
at that location.
!
! The static
forces acting!on the particle are Fg the gravity force, Fb the buoy !
ancy force, Fp the plasticity force, and three reaction forces FR1 ,
!
!
FR2 and FR3 , because the particle has three contact points with the
particles in the layer below. The dynamic forces induced by the
!
!
fluid movement are: FD the drag force and FL the lift force. The
particle can be dislodged from its position either by lifting (i.e., the
reactive forces become zero) or by rolling (i.e., the momentum of
the resulting forces relative to an axis of rotation defined by two of
the three contact points is greater than zero). The positions of the
points of contact are related to the separation distance between the
particles within a layer, and for the sake of simplicity, it will be
assumed that this distance is isotropic.
Choosing a coordinate system with axes dictated by the wellbore direction (Fig. 14), we can express the condition for lifting
as (note that the reaction forces are in this case zero)
! ! ! ! !
Fg Fb Fp FD FL 0 . . . . . . . . . . . . . . . . . 28
and the condition for rolling
!
!
!
!
!
~
t Fg ~
t  Fb ~
t  Fp ~
t  FD ~
t FL 0;
                   29
where ~
t is the vector between the axis of rotation (the line defined
by two of the three contact points) and the center of gravity and 
is the cross product of vectors. Note that the reaction force at the
opposite contact point to the rotation axis is zero when the rolling
starts and that the reaction forces at the contact points on the rotation axis do not contribute to the momentum because of the zero
length of the moment arm for those forces. The magnitude of each
force has been defined in different ways by several authors (Gavignet and Sobey 1989; Clark and Bickham 1994; Doron et al. 1997;
Kamp and Rivero 1999; Ramadan et al. 2001).
It is, however, difficult to evaluate the 3D fluid velocity around
a particle in the general case. By the use of statistical physic methods or CFD simulations as explained previously, it could be possible to derive a general law for the 3D fluid-velocity field as a
function of the flow rate, rotational velocity of the drillpipe, geometrical parameters of the system, and the rheology of the drilling
fluid, but that is an enormous task. We have adopted a simpler
strategy. Noticing that the preceding Eqs. 24 and 25 take a much
simpler form when there is no drillpipe rotation because we know
that the fluid-velocity field is parallel to the wellbore axis, we can
deduce the erosion rate of the top layer of the bed (noted l0Q , with
the dimension L2T1). This corresponds to the worst-case scenario
in which there is no assistance from drillpipe rotation. On the
other hand, the extended Larsens model can be used to find, for

March 2014 SPE Drilling & Completion


ID: jaganm Time: 16:20 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

13

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

Longitudinal view
Fb
FD

Fp

Fg

Page: 14

Total Pages: 17

Cross section view

Top view

vm
FL

FR1

Stage:

Possible axes if rotation

Axial component
of mud flow

FL

FR2

Fb
FD

Tangential
component
of mud flow

FR1

FR3

Fg

z
y

Fig. 14Forces acting on a single cuttings particle.

the given flow rate, the necessary rotational velocity (noted xcQ )
of the drillpipe that will ensure that particles cannot settle in a
bed. Then, an erosion rate l as a function of the drillpipe rotational velocity x can be defined as
lx

l0Q xcQ
; 8x < xcQ : . . . . . . . . . . . . . . . . . . 30
xcQ  x

If the used flow rate is insufficient to provoke any erosion of


the top layer of the cuttings bed, we search for the minimal flow
rate (Qc ) that causes either the lifting or rolling of the particles on
top of the cuttings bed while not rotating the drillpipes. With the
same procedure as described previously, an erosion function is
found for that flow rate. The generated erosion function is translated proportionally to the difference between the used flow rate
and the minimal flow rate to start erosion:
lx

l0Qc xcQc
B; 8x < xcQc  A; . . . . . . . . 31
xcQc  A  x

where the constants A and B are functions of (Q  Qc ) and shall


be chosen so that they are equal to zero when Q Qc .

A Designer Well in the North Sea


We will review two cases of poor cuttings-transport conditions in
a well recently drilled in the North Sea. This well was geometrically challenging and matched the definition given by Blikra et al.
(1994) of a designer well. According to this definition, a designer well is a well containing a combination of several of the
following elements:

Significant turn in the horizontal plane (e.g., 30 to 1508 azimuth) at high deviation
A combination of turnsboth right and left (i.e., snaky)
Turn not restricted by inclinations
The ability to place the borehole, as dictated by geology
The well that we will use for those two examples was
planned to be sidetracked from a main bore that was drilled in
the 90s. The sidetrack was planned to be drilled in two sections
(Fig. 15).
The 12 1/4 3 13 1/2-in. Section. The first section was a 12 1/4
 13 1/2-in. section. This section was kicked off from the main
bore with a previously installed whipstock in a sand-free formation. The whipstock was set in the 13 3/8-in. casing. A window
was milled through the 13 3/8-in. casing. This section was
planned to be drilled in one run with a rotary-steerable system and
a polycrystalline-diamond-compact (PDC) bit. An underreamer
was planned to be used to open up the hole to 13 1/2 in.
The well path started with a build at a rate of 3
/30 m to reach
an inclination of 658, then continued as a tangent to the section
total depth. WBM was used with a density of 1.55 specific gravity
(sg) at the beginning; it later increased to 1.62 sg toward the end
of the section. At the end of the drilling operation, a 10 3/4-in.
liner followed by a liner of 9 5/8 in. was set inside the 13 3/8-in.
casing.
The 8 1/2 3 9 1/2-in. Section. The second section was drilled
with an 8 1/2-in. PDC bit and a 9 1/2-in. underreamer. This was
the last section for this well, and it was drilled in one run with a
rotary-steerable system. In this section, the well path was steered
toward the targets with a maximal dogleg severity of 3
/30 m.

30 in. at 423 mMD


26 in. at 561 mMD

20 in. at 142 mMD


13 3 8 in. at 1783 mMD

10 in. at 2239 mMD


9 5 8 in. at 2247 mMD
8 in. at 2427 mMD
Fig. 15Trajectory and wellbore architecture of the test well.
14

March 2014 SPE Drilling & Completion


ID: jaganm Time: 16:20 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

8 in. bit

RSS

LWD

LWD

LWD

MWD

LWD

Stage:

Page: 15

Total Pages: 17

83 8 in. stab LWD 83 8 in. stab 5 in. NMDC HO and Reamer

5 in. HWDP

Jar

5 in. HWDP

5 in. DP

Fig. 16Layout of the 8 1/2 3 9 1/2-in. BHA.

Geosteering was used while drilling some intervals of this section.


The drilling fluid was an OBM with a density of 1.74 sg.

the cement plug. The cement drillout operation was performed


with flow rates of approximately 1900 L/min, 80-rev/min drillstring rotation, and an average drilling rate of 15 m/h.
Transport of different cuttings-particle sizes ranging from 2 to 6
mm was simulated to analyze the sensitivity of the results on the cuttings size in this particular operation. The results of the simulations
were matched with the observations made during the actual drilling
operations. From the simulations, it was clear that cuttings beds
formed regardless of the size of the particles but that the dimensions
(heights, lengths) of the modeled cuttings beds varied with the depth
position in the annulus and the cuttings-particle size. All the simulations resulted in no cuttings transport back to the surface with the
actual drilling parameters used in the drilling operation.
Further analysis showed that the drilling parameters used during
this operation were not sufficient to transport the cuttings particles.
The main factor that could have improved the transport capability
was the flow rate. However, it should be noted that the fluid velocity around the BHA was sufficient, and therefore created enough
lifting capability to transport the cuttings properly in that particular
area of the wellbore (Fig. 18). This is because of the very small
area between the BHA and the 9 5/8-in. liner. But when the flow
passed through a larger annular cross section, the fluid velocity
decreased, and the cuttings deposited in those sections of the well.
The larger the cross-sectional area in the wellbore, the greater the
height of the resulting cuttings bed. From the start of the cement
drillout operation, the lack of lifting capability caused cuttings beds
to accumulate in the inclined part of the wellbore.
The evolution of the cuttings bed during the drilling of the
cement plug is presented in Fig. 19 for different cuttings sizes.
The simulations pointed out that smaller particles were more difficult to transport than the larger ones. For small cuttings sizes, the
cuttings bed was thick even inside the 10 3/4-in. liner. A moreelongated and evenly distributed cuttings-bed shape was observed
when simulating with medium and large cuttings particles.
Close to the end of the cement drillout operation, a huge
amount of cuttings was accumulated in the inclined part of the
wellbore. This situation most likely led to a sudden cuttings avalanche, which is likely to be the cause of the major drilling packoff incident that occurred during this operation.

Summary of Operations. The cement plug was drilled, reaming


back every meter. The bottomhole assembly (BHA) was pulled
into the casing shoe, and an extended leakoff test (XLOT) was
performed. The formation was fractured during the XLOT, and a
cement stinger was run to set a new cement plug.
A second attempt was made to take the XLOT, but the BHA
got stuck in the fresh cement. The drillpipe was worked out and
circulated until the string was freed. At the third attempt, multiple
packoffs were experienced before reaching the formation. Eventually, the bottom of the cement plug was fractured, and mud losses
occurred. An XLOT was nevertheless performed, but the quality
of the results was uncertain; therefore, a new cement plug was set
and another XLOT taken, resulting in no doubt about the quality
of the measurements. It was decided to set another cement plug
and drill the section with MPD. Thereafter, the 8 1/2  9 1/2-in.
section was drilled successfully (Fig. 16).
First-Use Case: Poor Cuttings Transport While
Drilling a Cement Plug
Description of the Incident. During the third cement drillout
operation, there was a very poor cuttings-transport rate, which
was easily observed from the differences between the expected
active volume decrease and the real one (Fig. 17). At 1:00 pm the
active volume should have dropped by at least 5 m3, on the basis
of pure-cuttings removal, and more reasonably by 10 m3 if one
accounts for the expected loss of mud around the cuttings particles. There were some cuttings coming to the surface, but not
enough compared with the modeled well-cuttings return. Therefore, it was obvious that larger and larger cuttings beds were
forming in the borehole (Cayeux et al. 2012).
Analysis With the Transient Cuttings-Transport
Model
The aforementioned transient cuttings-transport model was used
to simulate the cuttings-transport behavior during the drilling of
22:00

0:00

2:00

Time

4:00

6:00

8:00

10:00

12:00

14:00
2250

2300 2350
Depth (m)

2400

10

20
30
ROP (m/h)

40

40

45 50 55 60
Active Volume (m3)

65

10
20
Cuttings flowrate (I/min)

Fig. 17This time-based log shows that the expected reduction of the active volume (rainbow-filled region), caused by cuttings removal, never happened (measured volume is displayed with blue solid line). One can also see that the expected cuttings-flow rate
at surface (green curve in cuttings-flow-rate track) is much higher than the actually measured cuttings-flow rate (blue curve in the
same track).
March 2014 SPE Drilling & Completion
ID: jaganm Time: 16:20 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

15

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

Fluid Velocity
200
400
600

Stage:

Total Pages: 17

Fluid density

Cuttings Proportion
0

Cuttings bed height

Page: 16

Because of small flow area between the bit/BHA and


the 9 5 8 in. liner, the fluid velocity just after the bit
was fast enough and therefore be able to transport
the cuttings upward in that particular area.

Temperature

200

200

400

400

600

600
800

1000

1000

1000

1000

1000

1200

1200

1200

1200

1200

MD

1400

1400

1400

1400

1600

1600

Depth (m)

800

Depth (m)

800

Depth (m)

800

MD (m)

800

1400
1600

1600

1600

1800

1800

1800

1800

1800

2000

2000

2000

2000

2000

2200

2200

2200

2200

2200

2400

2400

2400

2400

2400

2600

2600

2600

2600

2600

2800

2800

2800

2800
1.72 1.73 1.74
Density (sg)

2800

0
5
Fluid velocity (m/s)

Fluid velocity increase as the


cuttings bed height increases.
This is caused by the flow passes
through the smaller flow area
for a given flowrate.

0.1 0.2
Bed height

0
2
4
Cuttings proportion (%)

The cuttings proportion is showing the amount of


cuttings in suspension being transported. When the
flow area become smaller because of the cuttings bed, the
fluid velocity increases and therefore sufficent
enough to transport the cuttings properly upward.

0
20
40
Temperature (C)
Fluid density profile
accounted for the
influence of temperature
and the presence of
cuttings in the suspension.

Fig. 18Wellbore conditions after 9 hours drilling the cement plug.

Recommendations
Previous studies have shown that flow rate is the key parameter
affecting hole cleaning and cuttings-transport performance.
The use of high flow rate improves the material transport significantly, but at the same time it increases the risk of fracturing the formation as a result of high downhole pressure.
Because the cement drillout operation was performed inside a
closed annulus without any exposure of the drilling fluid to any
formation rocks, it would have been possible to increase the

flow rate well higher than 1900 L/min without risking formation
fracturing, at least until getting close to the end of the cement
plug.
Furthermore, increasing the drillstring rotation could have
improved the cuttings-transport performance. High drillstring rotation agitates the cuttings bed and therefore contributes to redistributing the cuttings bed into suspension. However, this would not
necessarily be a good option because the operation was carried out
while the BHA was inside a liner.

cuttings size = 2 mm

t+0

t+1

t+2

t+3

t+4

t+5

t+6

t+7

t+8

t+9 t+10 t+11 t+12 t+13 t+14 t+15

t+3

t+4

t+5

t+6

t+7

t+8

t+9 t+10 t+11 t+12 t+13 t+14 t+15

t+3

t+4

t+5

t+6

t+7

t+8

t+9 t+10 t+11 t+12 t+13 t+14 t+15

cuttings size = 4 mm

t+0

t+1

t+2

cuttings size = 6 mm

t+0

t+1

t+2

Fig. 19Evolution of cuttings-bed development during the cement drillout operation for different cuttings sizes.
16

March 2014 SPE Drilling & Completion


ID: jaganm Time: 16:20 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

Flow-rate (I/min)

Total Pages: 17

Flow-rate (I/min)

1 814

2 108

Hole cleaning index

Hole cleaning index

200

200

400

400

400

600

600

600

800

800

800

1000

1000

1000

1200

1200

1200

1400

1400

1600

MD

200

MD

MD

Page: 17

Flow-rate (I/min)

1 567
Hole cleaning index

Stage:

1600

1400
1600

1800
2000

1800
2000

1800
2000

2200

2200

2200

2400

2400

2400

2600

2600

2600

2800

2800

2800

1
0
0.5
Hole cleaning index

0
0.5
1
Hole cleaning index

0
0.5
1
Hole cleaning index

Fig. 20Effect of flow rate on hole cleaning and cuttings transport.

Reducing the drilling rate would have been another option to


improve the cuttings transport by counting on cuttings-bed erosion to transport the cuttings. Simulations made with the transient
cuttings-transport model show that the use of standard parameters
for the cuttings-bed erosion factors would have necessitated the
use of extremely low ROP to achieve that effect. This is because
the mud has a relatively high viscosity.
Second-Use Case: Hole-Cleaning Problems
During an MPD Operation
Indications of Poor Hole Cleaning. During the first days of the
MPD operation, indications of poor hole cleaning were visible.
The volume of cuttings that came back to the shakers was insufficient compared with the modeled cuttings returns. This was
observed by comparing the actual active volume variations with
the expected ones on the basis of the arrival time of cuttings to the
surface. At the same time, the downhole equivalent circulating
density (ECD) steadily increased during continuous periods of
drilling. However, at several instances, the drilling was interrupted for several hours to perform pore-pressure tests while
maintaining circulation. During those periods, in which no new
formations were drilled, the downhole ECD decreased to normal
values. After 4 days of such variable drilling conditions, it was
TD Speed (rpm)

60

Results From Transient Cuttings-Transport


Simulations
Three cuttings-size distributions were studied with the transient
cuttings-transport model using the real-time drilling parameters of
the MPD operation. The accumulation of cuttings in the regions
in which the internal diameter of the wellbore changed was confirmed by the simulations (liner hanger). As in the previous case,
cuttings beds developed in all studied cases. However, the shapes
and lengths of those cuttings beds were different from those in the
previous case. This was attributed to the use of a drilling fluid
with a much lower viscosity than in the cement drillout operation.
With a lower viscosity, it is easier to attain turbulence and therefore to increase the chance of redistributing the cuttings in the
cross-sectional area of the annulus for a better transport.
The effect of increasing the flow rate was seen in the simulations (Fig. 20). There was an improved transition in the holecleaning index as the flow rate was increased. Similarly, the effect
of increasing the pipe rotation was also seen in the simulations
(Fig. 21). Consequently, the geometry of the cuttings bed is
changed, as well as the hole-cleaning index.

Flow-rate (I/min)

TD Speed (rpm)

Flow-rate (I/min)

2 108

100

2 105

Cuttings proportion
0

Hole cleaning index

Cuttings bed height

TD Speed (rpm)

Cuttings proportion

Hole cleaning index

Cuttings bed height

Flow-rate (I/min)

2 098

21

Cuttings bed height

Hole cleaning index

0
200

400

400

400

400

600

600

600

600

600

600

600

600

600

800

800

800

800

800

800

800

800

800

1000

1000

1000

1000

1000

1000

1000

1000

1000

1200

1200

1200

1200

1200

1200

1200

1200

1200

1400

1400

1400

1400

1400
1600

1600

1600

1400
1600

1600

1600

1400
1600

MD (m)

400

MD (m)

400

200

Depth (m)

200

400

MD

200

400

MD (m)

200

400

Depth (m)

200

200

MD

200

200

Depth (m)

Depth (m)

Cuttings proportion
0

decided to reduce the ROP from 15 to 10 m/h. Thereafter, no


more problems with cuttings transport were noticed.

1400
1600

1400
1600

1800

1800

1800

1800

1800

1800

1800

1800

1800

2000

2000

2000

2000

2000

2000

2000

2000

2000

2200

2200

2200

2200

2200

2200

2200

2200

2200

2400

2400

2400

2400

2400

2400

2400

2400

2400

2600

2600

2600

2600

2600

2600

2600

2600

2600

2800

2800

2800

2800

2800

2800

2800

2800

0
5
10
Cuttings proportion (%)

0
30
Bed height (cm)

0
0.5
1
Hole cleaning index

0
5
10
Cuttings proportion (%)

0
30
Bed height (cm)

0
0.5
1
Hole cleaning index

0
5
10
Cuttings proportion (%)

2800
0
30
Bed height (cm)

1
0
0.5
Hole cleaning index

Fig. 21Effect of top-drive speed on hole cleaning and cuttings transport.


March 2014 SPE Drilling & Completion
ID: jaganm Time: 16:20 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

17

Bottom hole MD (m) ROP (m/h)

2 570,31

TD Speed (rpm) Flow-rate (I/min)

14,9

100

Stage:

Bottom hole MD (m) ROP (m/h)

2 067

2 570,30

Cuttings bed height

Cuttings proportion

15,0

Total Pages: 17

TD Speed (rpm)

2 066
Cuttings bed height

200

200

200

400

400

400

400

600

600

600

600

800

800

1000
1200

1000

1000

1200

1200

1200

1400
1600

1800
2000
2200

2200

20:30

1400
1600

MD (m)

800

1000
Depth (m)

1600

800

Time

1400

MD (m)

200

20:15

20:15

1400
1600

1800

1800

1800

2000

2000

2000

2200

2200

20:45

Flow-rate (I/min)

100

Cuttings proportion
0

Depth (m)

Page: 18

Time

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

2400

2400

2400

2400

2600

2600

2600

2600

2800

2800

2800

2800

20:45

21:00
0
5
10
Cuttings proportion (%)

10
Bed height (cm)

0
10
Cuttings flowrate

20:30

21:00
0
5
10
Cuttings proportion (%)

0
20
Bed height (cm)

0
60
Cuttings flowrate

0.0005

0.0001

Fig. 22Effect of erosion rate on cuttings transport.

The sensitivity of the simulations to the cuttings-beds erosion


factor (an internal calibration parameter of the model) was also
performed. Two simulation results are presented with a small factor and a factor five times larger (Fig. 22). The larger value of this
calibration factor gave results concordant with the observed cuttings-flow rate during the drilling operation. The estimation of the
actual cuttings-flow rate was performed by analyzing the rate of
change of the active pit volume during the operation when the
ROP was reduced to 10 m/h.
Analysis
At the start of the MPD operation, a flow rate of 1800 L/min was
used. After some time, it was increased to approximately 2100 L/
min, and there were improvements in relation to cuttings transport. From the simulations, it was seen that even at that flow rate,
hole cleaning was not efficient. Because we had an MPD operation in which pressure in the well was constantly monitored and
kept under control, it could have been possible to increase the
flow rate and compensate for the additional pressure losses in the
annulus by changing the settings of the MPD choke.
The pipe rotation was also critical. An increase in rotation
from 60 to 100 rev/min showed an improvementmost of the
cuttings in the beds were eroded and lifted up into the fluid flow.
The increase in rotation coupled with an increase in flow rate led
to the removal of the cuttings bed within the 10 3/4-in. liner.
The reduction of the ROP was a viable choice as seen from the
successful operation; however, it substantially increased the duration of the operation.
Conclusion
In the present work, a fully transient cuttings-transport model was
developed by integrating an empirical correlation for critical conditions of cuttings transport into a transient drilling model that accounts
for both fluid transport and drillstring mechanics. The model was calibrated against real data by adjusting parameters such as cuttings size
and cuttings-bed erosion factor with the measured cuttings rates at
surface as the best-fit criterion. A transient cuttings-transport model
makes it possible to better predict the downhole conditions because it
is able to represent phenomena that evolve over time, such as cuttings-bed buildup or removal. Effects related to changing operational
parameters are also taken into account so that the predicted well status
is as realistic as possible. With a real estimate of the downhole conditions, it is possible to provide operational recommendations that can
be used to avoid packoff and stuck-pipe incidents. By adjusting the
drilling parameters, such as the flow rate, drillstring rotational speed,
or the ROP, one can identify the adjustments that are needed to avoid
the formation of cuttings beds or the method to remove them.
18

The model has been used to analyze the cuttings-transport conditions for two situations taking place in a Designer Well in the
North Sea. In the first case, a cement drillout operation was analyzed. The operation took place in an inclined part of the well.
During the operation, several packoff incidents occurred that were
probably caused by cuttings-bed avalanches that led to a total
obstruction of the annulus. Simulations verified that, for the given
drilling parameters, cutting beds were evolving over time. It was
seen that the annular velocity was too small in parts of the well,
leading to a continuous buildup of a cuttings bed that most probably developed into a cuttings avalanche, which then resulted in
the subsequent packoff situations. The main operational recommendation was that a higher flow rate could have been used,
because there was no risk of fracturing the formation because the
operation took place within a cased hole.
In the second situation, an MPD operation was analyzed. Here,
the ROP was relatively high, leading to a situation in which cuttings were accumulating in the well. Simulations were performed
to reproduce the well conditions, and it was seen that the flow rate
used was too low. Because the well was operating in MPD conditions, there was an opportunity to increase the flow rate and to compensate for the increased downhole pressure by changing the
settings of the MPD choke. It was also confirmed by simulations
that an increase in rev/min would have been beneficial to decrease
the risk of cuttings accumulations by actively stirring the cuttings
beds. Simulations have also shown that material transport could
happen by cuttings-bed erosion, therefore justifying the reason that
the reduction of the ROP did stabilize the downhole conditions.
The simulations performed have confirmed that a transient cuttings-transport model is able to recreate the downhole well conditions, and that it can be a valuable tool for providing real-time
operational support and recommendations with respect to avoiding poor cuttings-transport conditions. Future work should focus
on performing more field studies to verify the applicability of the
model, as well as on improving the calibration of the model. Surface measurements of cuttings size and surface cuttings flow rate
could be integrated and linked more directly to the model in a
real-time environment for calibration purposes. Further model
improvements could also involve the inclusion of mud-gelling
effects and the incorporation of cuttings avalanches to make it
possible to predict potential packoff situations.
Nomenclature
A cross-sectional area
Aij NRTL parameters
Cd drag coefficient
Cpm specific heat capacity of the drilling fluid
March 2014 SPE Drilling & Completion

ID: jaganm Time: 16:21 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

Cpi specific heat capacity, with i w for water/


brine, o for oil, lgs for low-gravity solid, hgs for
high-gravity solid, c for cuttings, m for mud, l
for liquid, and s for solid
d diameter of the cuttings particles
di diameter of the cuttings particles of the ith size
!
Fb buoyancy force on a particle
!
FD drag force on a particle
!
Fg gravitational force on a particle
!
FL lift force on a particle
!
Fp plasticity force on a particle
! ! !
FR1 ; FR2 ; FR3 reaction forces on a particle
fc1 volume fraction of cuttings at initial conditions
of temperature T1 and pressure p1
fi volume fraction, with i w for water/brine, o
for oil, lgs for low-gravity solid, hgs for highgravity solid, c for cuttings, m for mud, l for liquid, and s for solid.
g gravitational acceleration of the Earth
H enthalpy per mass unit
K consistency index
ki thermal conductivity, with i w for water/brine,
o for oil, lgs for low-gravity solute, hgs for highgravity solid, c for cuttings, m for mud, l for liquid, and s for solid
L length of the controlled volume
m_ mass flux (per unit length) through the wall
NRe particle Reynolds number
n number of particles in the controlled volume
ni number of particles of the ith size in the controlled volume
p pressure
p1 reference pressure when cuttings are coming in
suspension
PEc packing efficiency of the coarse particles
PEi packing efficiency of the cuttings particles of
the ith size
PEf packing efficiency of the fine particles
PEmix packing efficiency of the mix of coarse and fine
particles
Qc conductive and natural-convective term
Qf forced convective term
qs heat generated by mechanical and hydraulic
frictions
r size ratio of the coarse particles to fine particles
R universal gas constant
s curvilinear abscissa
T temperature
T1 reference temperature when cuttings are coming
in suspension
Vi volume of phase i, with i w for water/brine, o
for oil, lgs for low-gravity solid, hgs for highgravity solid, c for cuttings, m for mud, l for liquid, and s for solid
!
va axial velocity of mud
vi slip velocity of the ith phase compared with vl
vl velocity of the liquid phase
vs cuttings-slip velocity
!
vt tangential velocity of mud
v!
m velocity of mud
Xf volume fraction of fine particles
a nonrandomness NRTL model
Cil interphase mass exchange
c_ shear rate
d separation between two particles
gmc viscosity of the drilling fluid with cuttings
gm viscosity of the drilling fluid
qi density, with i w for water/brine, o for oil, lgs
for low-gravity solid, hgs for high-gravity solid, c
for cuttings, m for mud, l for liquid, and s for solid

Stage:

Page: 19

Total Pages: 17

qm density of the drilling fluid


qm1 mud density at reference pressure p1 and temperature T1
fi mass fractions of the liquid phase, with i 2 H
l0Q cuttings erosion rate at current flow rate Q and
drillpipe rotation 0
k ratio of the solid thermal conductivity to the liquid thermal conductivity
xCQ critical rotational velocity to clear cuttings beds
at flow rate Q
X set of indices {w, o, lgs, hgs, c}
H set of indices {w, o}
W set of indices {lgs, hgs, c}
# average inclination
s shear stress
s0 yield stress

Acknowledgments
The presented work has been performed in the Centre for Drilling and Wells for Improved Recovery (SBBU). This is a cooperation between the International Research Institute of Stavanger
(IRIS), SINTEF, the University of Stavanger, and the Norwegian
University of Science and Technology (NTNU). SBBU is funded
by grants from the Research Council of Norway, Statoil, Total,
ConocoPhillips, Det Norske, Talisman Energy, and Wintershall.
The last author wants to thank Statoil for funding his position at
the University of Stavanger through the Akademia program.

References
Bassal, A.A. 1995. A Study of the Effect of Drill Pipe Rotation on Cuttings Transport in Inclined Wellbores. MS thesis, University of Tulsa,
Oklahoma.
Beaulne, M. and Mitsoulis, E. 1997. Creeping Motion of a Sphere in
Tubes Filled With HerschelBulkley Fluids. J. Non-Newtonian Fluid
Mechanics 72: 55-71.
Beirute, R.M. and Flumerfelt, R.W. 1977. An Evaluation of the Robertson-Stiff Model Describing Rheological Properties of Drilling Fluids
and Cement Slurries. SPE J. 17 (2): 97100. http://dx.doi.org/10.2118/
6505-PA.
Belarde, M. and Vestavik, O. 2011. Deployment of Reelwell Drilling
Method in Shale Gas Field in Canada. Paper SPE 145599 presented
at the SPE Offshore Europe Oil and Gas Conference and Exhibition,
Aberdeen, United Kingdom, 68 September. http://dx.doi.org/10.2118/
145599-MS.
Blikra, H., Drevdal, K.E., and Aarestad, T.V. 1994. Extended Reach, Horizontal and Complex Wells: Challenges, Achievements and Cost-Benefits. Paper SPE 28005 presented at the University of Tulsa Centennial
Petroleum Engineering Symposium, Tulsa, Oklahoma, 2931 August.
http://dx.doi.org/10.2118/28005-MS.
Cayeux, E., Daireaux, B., Dvergsnes, E. et al. 2012. An Early Warning
System for Identifying Drilling Problems: An Example From a Problematic Drill-Out Cement Operation in the North-Sea. Paper SPE
15942 presented at the SPE Drilling Conference in San Diego, California, 68 March. http://dx.doi.org/10.2118/15942-MS.
Cayeux, E. and Lande, H.P. 2013. Factors Influencing the Estimation of
Downhole Pressure far Away From Measurement Points During Drilling Operations. Paper presented at the SIMS 2013 conference in Bergen, Norway, 1618 October.
Clark, R.K. and Bickham, K.L. 1994. A Mechanistic Model for Cuttings
Transport. Paper SPE 28306 presented at the SPE Annual Technical
Conference and Exhibition, New Orleans, Louisiana, 2528 September. http://dx.doi.org/10.2118/28306-MS.
DDBST. 2014. Dortmund Data Bank Software & Separation Technology
GmbH. http://www.ddbst.com/ddb-search.html (accessed 19 February
2014).
Doron, P., Simkhis, M., and Barnea, D. 1997. Flow of Solid-Liquid Mixtures in Inclined Pipes. International J. Multiphase Flow 23 (2):
313323.

March 2014 SPE Drilling & Completion


ID: jaganm Time: 16:21 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

19

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

Einstein, A. 1906. Eine Neue Bestimmung Der Molekuldimensionen.


Annalen Der Physik 19 (5).
Eucken, A. 1940. Allgemeine Gesetzmassigkeiten fur das Warmeleitvermogen verschiedener Stoffarten und Aggregatzustande. Forschung auf
dem Gebiete des Ingenieurwesens. 11 (1): 620. ISSN 00157899.
Ford, J.T., Peden, J.M., Oyeneyin, M.B. et al. 1990. Experimental Investigation of Drilled Cuttings Transport in Inclined Boreholes. Paper SPE
20421 presented at the Annual Technical Conference and Exhibition
of the SPE, New Orleans, Louisiana, 2326 September. http://
dx.doi.org/10.2118/20421-MS.
Gauss, C.F. 1831. Besprechung des Buchs von L.A. Seeber: Intersuchungen uber die Eigenschaften der positiven ternaren quadratischen Formen usw. Gottingsche Gelehrte Anzeigen.
Gavignet, A.A. and Sobey, I.J. 1989. Model Aids Cuttings Transport Prediction. J. Pet Tech 41 (9): 916921. http://dx.doi.org/10.2118/15417-PA.
Gmehling, J. and Onken, U. 1977. Vapour-Liquid-Equilibrium Data Collection, Vol. 1, Part 2a, Frankfurt: Verlag & Druckerel Friedrich
Bischoff.
Hareland, G., Azar, J.J., and Rampersad, P.R. 1993. Comparison of Cuttings Transport in Directional Drilling Using Low-Toxicity Invert
Emulsion Mineral-Oil-Based and Water-Based Muds. Paper SPE
25871 presented at the Low Permeability Reservoirs Symposium,
Denver, Colorado, 2628 April. http://dx.doi.org/10.2118/25871MS.
Hastchek, E. 1910. Die Viskositat Der Dispersoide. Kolloid-Zeitschrift 8:
3439.
Holtzer, A. and Sommerfeld, M. 2008. New Simple Correlation Formula
for the Drag Coefficient of Non-Spherical Particles. Powder Technol.
184: 361365.
Houwen, O.H. and Geehan T. 1986. Rheology of Oil-Based Muds. Paper
SPE 15416 presented at the SPE Annual Technical Conference, New
Orleans, Louisiana, 58 October. http://dx.doi.org/10.2118/15416-MS.
Isambourg, P., Anfinsen, B.T., and Marken, C. 1996. Volumetric Behavior
of Drilling Muds at High Pressure and High Temperature. Paper SPE
36830 presented at the European Petroleum Conference, Milan, Italy,
2224 October. http://dx.doi.org/10.2118/36830-MS.
Jalukar, L.S. 1993. Study of Hole Size Effect on Critical and Subcritical
Drilling Fluid Velocities in Cuttings Transport for Inclined Wellbores.
MS thesis, University of Tulsa, Oklahoma.
Kamp, A.M. and Rivero, M. 1999. Layer Modeling for Cuttings Transport
in Highly Inclined Wellbores. Paper SPE 53942 presented at the SPE
Annul Technical Conference and Exhibition, Caracas, Venezuela,
2123 April. http://dx.doi.org/10.2118/53942-MS.
Kandula, M. 2011. On the Effective Thermal Conductivity of Porous
Packed Beds With Uniform Spherical Particles. J. Porous Media 14
(10): 919926.
Kemp, N.P., Thomas, D.C., Atkinson, G. et al. 1989. Density Modeling
for Brines as a Function of Composition, Temperature, and Pressure.
SPE Res Eng 4 (4): 394400. http://dx.doi.org/10.2118/16079-PA.
Lage, A., Fjelde, K.K., and Time, R. 2003. Underbalanced Drilling Dynamics: Two-Phase Flow Modeling and Experiments. SPE J. 8 (1):
6170. http://dx.doi.org/10.2118/83607-PA.
Larsen, T.I., Pilehvari, A.A., and Azar, J.J. 1997. Development of a New
Cuttings-Transport Model for High-Angle Wellbores Including Horizontal Wells. SPE Drill & Compl 12 (2): 129136. http://dx.doi.org/
10.2118/25872-PA.
Liles, D.R. and Reed, W.H. 1978. A Semi-Implicit Method for Two-Phase
Dynamics J. Computational Physics 26.
Loth, E. 2008. Drag of Non-Spherical Solid Particles of Regular and Irregular Shape. Powder Technol. 182: 342-353.
Marshall, D.W. and Bentsen R.G. 1982. A Computer Model to Determine the
Temperature Distributions in a Wellbore. J. Cdn. Pet. Tech. 21 (1): 63-75.
Maxwell, J.C. 1873. Treatise on Electricity and Magnetism. 1. Clarendom
Press. Oxford, U.K.
Nagata, I. 1973. Prediction Accuracy of Multicomponent Vapour-Liquid
Equilibrium Data From Binary Parameters. J. Chem. Eng. Japan 6 (1).
Nazari, T, Hareland, G., and Azar, J.J. 2010. Review of Cuttings Transport
in Directional Well Drilling: Systematic Approach. Paper SPE 132372
presented at the SPE Western Regional Meeting, Anaheim, California,
2729 May. http://dx.doi.org/10.2118/132372-MS.
Ozbayoglu, M.E., Saasen, A., and Sorgun, M. 2008. The Effect of Pipe
Rotation on Hole Cleaning for Water-Based Drilling Fluid in Hori20

Stage:

Page: 20

Total Pages: 17

zontal and Deviated Wells. Paper SPE 114965 presented at the


IADC/SPE Asia Pacific Drilling Technology Conference and Exhibition, Jakarta, Indonesia, 2527 August. http://dx.doi.org/10.2118/
114965-MS.
Peden, J.M., Ford, J.T., and Oyeneyin, M.B. 1990. Comprehensive Experimental Investigation of Drilled Cuttings Transport in Inclined Wells
Including the Effects of Rotation and Eccentricity. Paper SPE 20925
presented at Europec, The Hague, Netherlands, 2224 October. http://
dx.doi.org/10.2118/20925-MS.
Pilehvari, A.A., Azar, J.J., and Shirazi, S.A. 1999. State-of-the-ArtCuttings Transport in Horizontal Wellbores. SPE Drill & Compl 14 (3).
http://dx.doi.org/10.2118/57716-PA.
Ramadan, A., Skalle, P., Johansen, S.T. et al. 2001. Mechanistic Model
for Cuttings Removal From Solid Bed in Inclined Channels. J. Pet Sci
& Eng 30 (34): 129141.
Ranjbar, R. 2010. Cuttings Transport in Inclined and Horizontal Wellbore.
MS thesis, University of Stavanger, Norway (June 2010).
Renaud, M., Mauret, E., and Chhabra, R. 2004. Power-Law Fluid Flow
Over a Sphere: Average Shear Rate and Drag Coefficient. Canadian J.
Chem. Eng. 82: 1066-1070.
Robertson, R.E. and Stiff, H.A. 1976. An Improved Mathematical
Model for Relating Shear Stress to Shear Rate in Drilling Fluids and
Cement Slurries. SPE J. 16 (1): 3136. http://dx.doi.org/10.2118/
5333-PA.
Rowley, R. 1981. A local Composition Model for Multicomponent Liquid Mixture Thermal Conductivities. Chem. Eng. Sci. 37 (6):
897904.
Rubiandini, R.S. 1999. Equation for Estimating Mud Minimum Rate for
Cuttings Transport in an Inclined-Until-Horizontal Well. Paper SPE
57541 presented at the SPE Annual Technical Conference and Exhibition, Abu Dhabi, UAE, 810 November. http://dx.doi.org/10.2118/
57541-MS.
Song, C., Wang, P, and Makse, H.A. 2008. A Phase Diagram for Jammed
Matter. Nature 453: 629-632.
Tomren, P.H., Iyodo, A.W., and Azar, J.J. 1986. Experimental Study of
Cuttings Transport in Directional Wells. SPE Drill Eng 1 (1): 4356.
http://dx.doi.org/10.2118/12123-PA.
Trapp, J.A. and Riemke, R.A. 1986. A Nearly-Implicit Hydrodynamic
Numerical Scheme for Two-Phase Flows. J. Computational Physics
66.
Whittaker, A. 1985. Theory and Application of Drilling Fluid Hydraulics.
Vol. 1, Book series: Exlog Series of Petroleum Geology and Engineering Handbooks, Springer. ISBN 0-88746-045-3.
Yao, D. and Robello, S. 2008. Annular Pressure Loss Predictions for Various Stand-off Devices. Paper SPE 112544 presented at the SPE/IADC
Drilling Conference, Orlando, Florida, 46 March. http://dx.doi.org/
10.2118/112544-MS.
Zheng, J., Carlson, W., and Reed, J. 1995. The Packing Density of
Binary Powder Mixtures. J. European Ceramic Society 15 (5):
479483.

Eric Cayeux currently holds the position of chief scientist at IRIS


in Norway. He joined IRIS in 2004, following a 20-year career in
the oil and gas industry, specializing in drilling and well-software solutions. Cayeux earned an MSc degree in civil engineering from the Ecole Nationale des Travaux Publics de
lEtat, Lyon, France, and an MSc degree in software engineering from the University of Nice in association with the Paris
School of Mines, France. He is an SPE member.
Taiwo Mesagan currently works with Statoil in Norway as a drilling and well engineer since he joined in 2012. He earned a
BSc degree in petroleum and gas engineering from the University of Lagos, Nigeria, before studying further for an MSc
degree in petroleum engineering from the University of Stavanger. Mesagan is an SPE member.
Sakti Tanripada started his career in the oil industry as a petroleum engineer for Chevron. He then continued as a drilling
and completion engineer for Statoil after earning an MSc
degree in petroleum engineering from the University of Stavanger. Tanripada also earned a BSc degree within the same
subject from the Bandung Institute of Technology, Indonesia.
He is an SPE member.
March 2014 SPE Drilling & Completion

ID: jaganm Time: 16:21 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

DC163492 DOI: 10.2118/163492-PA Date: 15-March-14

Mohamed Zidan currently holds the position of principal engineer at Statoil ASA in Norway. He joined Statoil in 2009, following a 9-year career in the oil and gas industry, specializing in
measurement while drilling/logging while drilling, directional
drilling, drilling engineering, and implementation of new drilling technologies. Zidan earned a BSc degree in industrial and
systems engineering from North Carolina A&T State University,
USA. He is an SPE member.

Stage:

Page: 21

Total Pages: 17

Kjell Ka
re Fjelde is currently working as a professor in drilling engineering at the Department of Petroleum Engineering, University
of Stavanger. His main research areas include well-flow modeling and well control. Fjelde previously worked as a teacher at
Stavanger Offshore Technical College and as a research scientist and group leader at IRIS. He holds a PhD degree in applied
mathematics, which focused on numerical modeling of multiphase flow, from the University of Bergen, Norway.

March 2014 SPE Drilling & Completion


ID: jaganm Time: 16:21 I Path: S:/3B2/DC##/Vol00000/140005/APPFile/SA-DC##140005

21

Das könnte Ihnen auch gefallen