Вы находитесь на странице: 1из 38

SPE-170330-MS

Principles and Sensitivity Analysis of Automatic Calibration of MPD


Methods Based on Dual-Gradient Drilling Solutions
E. Cayeux, B. Daireaux, E.W. Dvergsnes, H. Siahaan, and J.E. Gravdal, IRIS

Copyright 2014, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Deepwater Drilling and Completions Conference held in Galveston, Texas, USA, 10 –11 September 2014.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents
of the paper have not been reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect
any position of the Society of Petroleum Engineers, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written
consent of the Society of Petroleum Engineers is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may
not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
Dual-gradient drilling (DGD) solutions have been developed to work with greater margins while drilling
in deep water environments by mirroring hydrostatic pressure conditions that are closer to the natural
ones. Recent attempts have been undertaken to use DGD to actively control the downhole pressure and
therefore turning them into managed pressure drilling (MPD) solutions. The pressure control strategy
requires the knowledge of how pressure will change along the borehole as a function of the drilling
parameters and therefore needs to use numerical hydraulic models. However, the accuracy of the
prediction of pressure calculations depends on a certain number of parameters that are not necessarily well
known. Some of those significant factors are well defined but are not necessarily measured as extensively
as it would have been required. Others are subject to the actual downhole conditions and may change with
time. Therefore, they are, in many instances, difficult to assess with enough certainty. As a result, it is
necessary to constantly calibrate the downhole pressure calculations in order to match observed values at
controlled positions. This calibration process is usually done manually by a human operator, leaving the
possibility for possible misinterpretations of the actual calibration data and consequently a potentially
erroneous control of the downhole pressure by the MPD control algorithm.
In this paper, we review the sources of uncertainties that can affect the accuracy of the downhole
pressure calculations. Thereafter, we explain how automatic calibration of the numerical models can be
made in order to match reference measurements. The proposed method also allows for an evaluation of
the accuracy by which the downhole pressure calculations can be made. In conjunction with the required
precision to control the downhole pressure during a drilling operation, it is possible to assess whether
additional measurements or working procedures should be implemented prior to an MPD dual-gradient
operation with tight constraints.

Introduction
With the availability of higher generations of semi-submersible rigs, it has been possible to explore and
develop oil and gas fields in greater water-depths. However, there are additional challenges with the
geo-pressure margins when the seabed depth increases.
The overburden pressure, also called lithostatic pressure, is defined as:
2 SPE-170330-MS

Figure 1—Conventional drilling program in deep water drilling.

(1)

where TVD is the true vertical depth of the point of interest, TVD0 is the true vertical depth at origin,
pl is the lithostatic pressure, ␳b is the bulk density of the material at a particular depth and g is the
gravitational acceleration. In an offshore context, the latter equation can be re-written (Aadnøy 1998):
(2)

where TVDseabed is the vertical depth of the seabed and ␳w is the water density. The density of water
changes slightly with temperature and pressure, but insignificantly compared to the change in bulk density
of formation rocks as a function of depth, simply because of the increased effect of compaction on rock
material at increasing depths.
As a consequence, there is clearly an important role played by the water depth on the gradient build-up
of the lithostatic pressure as a function of depth. Since the pore, collapse and fracturing pressures
originates from the sea bottom, the use of a single mud weight into the annulus does not fit well with the
formation geo-pressure margins (see Figure 1). As a consequence, the pressure window only allows
relatively short sections with a given mud weight and therefore it is necessary to use many casing strings
to reach the total depth (TD) of the well.
To get a better match, the mud column inside the well should mimic the natural ground conditions,
which means using two mud weights (see Figure 2).
There are several solutions to obtain the dual mud weight in the annulus. One consists of injecting light
weight additives (e.g. gas, glass beads) into the mud contained in the marine riser in order to reduce the
apparent mud density above the sea level (Kozicz et al. 2006). Alternatively, subsea pumping can be used
to lift up the mud to the rig while keeping a static column of another fluid of different density, sometimes
called the blanket fluid, above the suction point.
If the blanket fluid is sea-water and its interface level shall not change, then it is actually not necessary
anymore to have a marine riser, therefore resulting in a riser-less drilling solution (Stave et al. 2005).
However, if one wants to actively control the downhole pressure at a given depth of the borehole, it is
possible to vary the level of the interface between the blanket fluid and the drilling fluid in the annulus,
thus changing the hydrostatic component of the downhole pressure. In that case, the dual-gradient method
turns into an MPD solution. To achieve this goal, it is not only necessary to use a lift pump to divert the
drilling fluid from the annulus before it invades completely the marine riser, it is also necessary to refill
the annulus with blanket fluid if the interface level shall be lowered. In addition, in order to assist in
SPE-170330-MS 3

Figure 2—Drilling program based on a dual gradient mud column.

Figure 3—Dual-gradient drilling

raising the interface level, it is useful to use a booster pump that can inject drilling fluid just below the
lift pump inlet (see Figure 3).
The fact that the annulus is filled with two fluids of different densities, while the drill-string contains
a single fluid, i.e. the drilling mud, causes a problem because the hydrostatic head of the fluid in the
drill-string is different from the one in the annulus. To manage that problem, either one has to let the level
of the fluid in the drill-string drop when the mud pumps are stopped until reaching an equilibrium between
the two branches of the equivalent U-tube, or one shall use a special valve, a so-called drill-string valve
(DSV), into the drill-string that closes when the circulation is stopped (Goldsmith 1998). Regardless of
the chosen solution, the imbalance between the two branches of the U-tube, generates challenges during
well control situations (Schubert et al. 2006).
As a special type of blanket fluid, air can be used instead of water. In that particular case, the dual
gradient method is called a low annulus level return method. The advantage of a good match between the
4 SPE-170330-MS

Figure 4 —Inherent U-tube effect of a dual gradient method: the drill-string is filled with a drilling fluid while the annulus is filled with drilling fluid
at the bottom and a blanket fluid at the top.

Figure 5—Low-level annulus return drilling

formation pressure gradient and the borehole pressure gradient is partially lost but other advantages can
be gained. One of them is a simplification of the top side equipment since it is not necessary anymore to
install a pumping mechanism to refill the annulus if the interface level between the blanket fluid and the
drilling fluid should be lowered (see Figure 5). It is also possible to only lower the interface level during
circulation time in order to compensate for the pressure losses due to the drilling fluid movement (Ziegler
et al. 2013). When circulation stops then the drilling fluid level raises back to the level of the outlet of
the well as it would have been in a conventional drilling method, resulting in a natural hydrostatic
overbalance and therefore eliminating the possible well control problems mentioned above.
Whether a deep water well is drilled conventionally or by using a dual gradient method, the cold marine
environment influences strongly the heat exchange. In turn, the cold temperatures have large effects on
the mud characteristics, e.g. density, rheology, and therefore influences the downhole pressure both in
hydrostatic and hydrodynamic conditions. Drilling in deep water is therefore challenging and control
methods could aid leveraging the execution of such drilling operations. However, in order to achieve an
acceptable control of the downhole pressure, it is necessary to evaluate the uncertainty of the model
estimations that are made in the process control and at the same time to attempt at calibrating the
calculations with the partial measurements that are available.
SPE-170330-MS 5

Hydraulic Model

Equations for the hydraulic system


The behavior of the drilling fluid is governed by three balance equations that describe the interface
exchange of mass (Lavoisier) and momentum (Navier-Stokes) and energy (based on the first law of
thermodynamics).
Drilling fluids are often a mix of different components. Some of those components are part of the fluid
formulation, while others result from the drilling process itself like cuttings and formation fluids. The
components of the fluid can be found in different phase states, including liquid phases like brine and oil,
solid phases as for example bentonite, barite, lost circulation materials (LCM), cuttings and gaseous
phases as for instance formation gas. As a consequence, multi-phase and multi-component flow shall be
used. It should be noted that in normal drilling conditions, the dimensions of the solid particles, liquid
droplets and gas bubbles remain very small compared to the dimensions of a cross-section of the conduits
where the fluid flows. As a consequence, it is acceptable to use a drift-flux formulation where the different
phases are mixed together but each component has a slip velocity with regard to a reference one. The
drift-flux approximation allows to transform the original 3-dimensional problem into a 1-dimension one,
thus simplifying greatly the calculations.
With the drift-flux approximation, the mass balance for each component can be written (see derivation
in Appendix A)(Frøyen and Sævareid 2000):
(3)

where i is an index representing a component of the fluid, ⍀ is a set of components, t is time, s is the
curvilinear abscissa, p is the pressure, A is the cross-sectional area of a fluid element, is the
mixture compressibility factor, ␳mix is the density of the mix, F is the total mass flux, ṁi is the mass flux
per unit length through the walls of component i.
Similarly, the momentum balance can be written as follow (see derivation in Appendix B)(Frøyen and
Sævareid 2000):
(4)

where is the centre of mass velocity, ␮ is the shear stress function,


with being the cross-sectional average velocity of a phase k and xk being its
mass fraction, ␽ is the average inclination of the fluid element and is the velocity of the mass flux
through the walls.
The conservation of energy for a drilling fluid in a wellbore can be as formulated by Corre et al. (1984):
(5)

where H is enthalpy and qs is the heat generation in the system caused by mechanical (rotational)
energy and hydraulic energy. Qf is the forced convective term which may be expressed by:
(6)

Further, Qc is the conductive and natural conductive term that does not have a general expression. In
the case of purely convective isotropic material Qc may be expressed as:
(7)

where ␭ is thermal conductivity and T is temperature.


6 SPE-170330-MS

The friction pressure loss in equation (4), expressed by ␮, depends on the rheology of the drilling fluid
and the flow regime (laminar, transitional or turbulent). Drilling muds are shear thinning fluids with yield
stress which rheology can be approximated by three parameters constitutive laws like Herschel-Bulkley:
(8)

where ␶ is the shear stress, ␶0 is the yield stress, k is the consistency index, is the shear rate and n
the flow behavior index, or the one of Robertson and Stiff (1976):
(9)

where A, B and C are the coefficients of the model. The expression of the friction pressure loss for both
Herschel-Bulkey’s and Robertson-Stiff’s constitutive laws in pipe and annular can be found, for example,
in Ochoa (2006).
Note that the solution to the system of partial differential equations defined by (3), (4) and (5) depends
on the boundary conditions which may change through time. It is therefore natural to integrate the system
of partial differential equations using a stepwise integration in time. The stability of the integration of such
hyperbolic partial differential equations is conditioned by the Courant–Friedrichs–Lewy (CFL) criteria
(Courant et al. 1967) and because of the very high velocity of pressure waves in drilling fluids, any
explicit method may require very small time steps to be stable. By considering the pressure wave
propagation implicitly, the focus is then shifted to the evolving process of mass redistribution which is
much slower than the one of acoustic wave propagations. Therefore, reasonable time steps can be used in
order to respect the CFL condition. After discretization of those equations in space, a semi-implicit finite
difference method is used to solve the mass transport and pressure distribution (Liles and Reed 1978,
Trapp and Riemke 1986) in a continuous conduit, possibly of variable size.

Hydraulic network
Discontinuities, like concentrated sources (e.g. pumps), abrupt change of cross-sectional area (e.g. bit,
valves) or junctions where three pipes meet together, delimit the start and end of hydraulic branches. The
connection between the different branches defines a hydraulic network. The simplest hydraulic network
can be found with conventional drilling. It has two branches with the drill-pipe on one side connected
through the bit to the annulus on the other side. The dual-gradient drilling methods are necessarily
described by a hydraulic network with more than two branches since the annulus has at least a junction
where the lift pump is connected. In addition, downhole equipment may create discontinuities like valves
or connect the interior of the drill-string with the annulus at other places than the bit. Figure 6 shows and
example of such a network for a complete dual-gradient solution. Note that there are possible connections
between the drill-string interior and the annulus at the level of the circulation sub, the under-reamer and
the downhole motor (PDM), therefore creating even more branches in the hydraulic circuit.
For area discontinuities, there is a pressure jump between the end of the one branch and the start of the
other branch. A pump is described by a source term in the form of a mass flux. Junctions are characterized
by equal pressure at the connection between the three branches and an equilibrium of the mass fluxes (the
sum of the mass fluxes of at the junction is 0). Therefore a junction is characterized by four unknowns:
three mass fluxes and one pressure. A junction system of dimension 4m ⫻ 4m, where m is the number of
junctions, can be assembled and solved, therefore providing the missing boundary conditions for all the
branches of the hydraulic network.
The described hydraulic model can calculate the transient response of a complex hydraulic network
with multiple fluids as a function of continuously changing boundary conditions and therefore is well
suited for working with control problems in a dual-gradient drilling context.
SPE-170330-MS 7

U-tube Effect
In conventional drilling operations it is usually of
interest to pull a dry string when pulling out of hole
to avoid spilling drilling fluid on the drill-floor. To
accomplish this, the U-tube effect is exploited by
putting a dense slug pill in the drill-pipe. The U-tube
effect will then balance the pressure from the mud
inside the drill-string by the pressure caused by the
mud inside the annulus such that there will be a
non-zero liquid level (air above the slug pill) inside
the drill-string. Note that this effect is also used in
other operations such as well control situations and
during cementing.
In the dual gradient drilling solutions discussed
here, there is always a U-tube effect present when-
ever the main pumps are stopped, for example dur- Figure 6 —Example of a hydraulic network that represents a dual-
ing connections. The reason is that hydrostatic pres- gradient drilling method with some special elements in the drill-string/
BHA that connect the drill-string interior with the annulus.
sure from the riser, which is partially filled with air
or blanket fluid, cannot balance the hydrostatic
pressure from the drilling fluid inside the drill-string. Therefore, the drilling fluid continues to flow out
of the drill-string and into the annulus until a sufficient liquid level is obtained inside the drill-string. This
means that the lift pump has to continue pumping until the flow caused by the U-tube effect stops. An
alternative to avoid this situation is to put a DSV in the BHA to block the U-tube effect.
The above described hydraulic model can calculate the transient behavior that takes place during the
equalization of the pressures between the drill-string and the annulus, yet accounting for variable
boundary conditions that are imposed by any control algorithm of the lift pump flow-rate.
To illustrate the U-tube effect, we will use an example based on a dual-gradient riser-less drilling
solution in 1000 m water depth. Because of the riser-less constraint, the control strategy of the lift pump
flow-rate is to maintain constant the interface level between the top of the drilling mud column in the
annulus and the sea-water. A very accurate pressure sensor placed below this interface level is used to
estimate its elevation. We will refer to that sensor as the reference pressure sensor. A simple PI
(Proportional Integral) controller adjusts the lift pump flow-rate, i.e. the process variable of the feed-back
control algorithm, in order to keep constant the reference pressure, i.e. the set-point of the PI controller.
The vertical section and horizontal projection of the trajectory used in the following calculations are
shown on the lower left-hand side of Figure 7 (c). For the following calculations, we will focus on the
drilling of the 8 ½-in hole size using a SBM of density 1.67 s.g. The 9 5/8-in casing shoe is set at
7500mMD.
The first calculations are made at bit depth 7600 mMD, when the main mud pumps, which were
running at 2000 l/min, are stopped. Note that the flow-rate of the booster pump is kept constant at 2000
l/min during the whole period. On the top left hand-side of Figure 7 (a), we can see that the lift pump
flow-rate reduces while the flow-rate of the well decreases due to the stopping of the mud pumps. But
when the liquid level in the drill-string starts to drop, the flow-rate induced by the U-tube effect picks up
in intensity and the lift pump flow-rate adjusts itself to maintain a constant interface level between the
drilling mud in the annulus and the sea water. Note that the flow from the drill-pipe into the annulus
reduces quickly during the first 12 minutes (see Figure 7 (b)), and thereafter it is just a few ten liter per
minute. The flow rate from the drill-string to the annulus is still 5 l/min one hour after the flow reduction
at the main pump, and the liquid level in the string did not settle completely until almost 2 hours and 20
minutes after the flow reduction at the main pump.
8 SPE-170330-MS

Figure 7—U-tube effect when the mud pumps are stopped in a riser-less dual gradient drilling method in 1000m water depth with a 8 ½-in BHA at
bit depth @ 7600mMD.

Compressibility
The compressibility of the drilling fluid plays a role in the mass transfer described by eq. (3). In practice,
it means that when the pressure increases, a larger quantity of fluid is required to fill the same volume.
The compressibility of water and oil base mud is relatively small, but for a large volume and for a wide
range of pressure, it may nevertheless result in a relatively large quantity of additional fluid that needs to
be pumped into the well.
This is illustrated by the example of Figure 8. In this riser-less dual-gradient setup without any DSV
in the BHA, circulation is established at 200 l/min. Because of the U-tube effect, the initial liquid level
in the drill-string is at 300 meter before the start of the mud pumps. During the first 50 seconds, the fluid
velocity in the drill-string increases gradually in depth and magnitude until the fluid starts to move on the
lower side of the annulus. Until this moment, the downhole pressure at the bit has not changed yet because
the increase in pressure only concerns the fluid contained in the drill-string. When the fluid in the annulus
starts to move, the velocity of the fluid in the drill-string first reduces because part of the pressure that had
built up in the drill-string is evacuated in the annulus. However, afterward the fluid in the drill-string
regain force while more and more of the fluid contained in the annulus starts to move. Because of the
larger volume, it takes longer time to move the entire annulus mud column and it is only after 180 seconds
that, finally, the fluid starts to move just below the lift pump inlet. At that moment, the pressure increases
at the level of the reference pressure and the lift pump flow-rate starts to accelerate. It takes another 100
seconds before the flow-rate out equalizes with the flow-rate in. Note that during the whole sequence,
there is in fact no main mud pump pressure because the whole circulation is caused by the U-tube effect.
SPE-170330-MS 9

Figure 8 —Circulation is established at 200 l/min in a riser-less dual gradient setup without the use of a DSV. The initial liquid level in the drill-string
is 300 m corresponding to the U-tube natural equilibrium.

In this particular case, it would require a main mud pump flow-rate larger than 900 l/min before the liquid
level in the drill-string reaches the top of the drill-string resulting in a measurable pump pressure.
This minimum flow rate to compensate for the U-tube effect is of course dependent on the difference
between the hydrostatic pressure of the mud column inside the drill-string (above the control point)
compared to the pressure at the control point in the annulus. In the riser-less dual-gradient drilling solution
discussed here, the minimum flow rate will therefore depend on the mud density and the water depth.
According to these simulations, it would take as long as 350 seconds before reaching an equalization
of the flow-rate out with the flow-rate in, when pumping at 200 l/min in those conditions. This may seem
surprising but in fact, the accuracy of the calculations made with the described transient model can be
verified in a real drilling operation performed in similar conditions, although using conventional drilling
and not dual-gradient drilling methods.
This example (shown on Figure 9) is for a reaming operation performed at bit depth 8100 mMD. Each
time the mud pumps were started, it took more than five minutes before the flow-rate out reached the same
level as the flow-rate in. Similarly, when the mud pumps were stopped, it took about two minutes before
the mud flow-rate out stopped. It should be noted that, at this rig site, the paddle did not indicate any
values unless the flow-rate was above 500 l/min. Despite this threshold, the graphs show a good agreement
between the amplitude and timing of the calculated flow-rate out and the paddle measurements, at least
when the flow-rate out was greater than 500l/min. The additional flow-out resulting from the downward
movement of the drill-string is also clearly visible on the paddle readings and on the calculated flow-rate
out.
10 SPE-170330-MS

Figure 9 —Comparison between calculated and measured values when reaming down one stand in a long horizontal well (bit depth 8080mMD).

Dependence on Temperature
The density and viscosity of the drilling fluid at any depths depend on the local pressure and temperature
conditions. Therefore the solution closure depends on the estimation of the downhole temperature at any
depth of the well.
As for downhole pressure, downhole temperature measurements are very sparse, if available at all. It
is possible to calculate the temperature of the drilling fluid at any depth as a function of time by using a
transient heat transfer model (Marshall and Bentsen 1982, Corre et al. 1984) as shown in equation (5). It
should be noted that the heat transfer process never reaches any steady state conditions because the mud
pumps are stopped and started each time a new stand is added, and because the depth of the well
continuously changes due to the drilling process, hence exposing deeper formations at higher temperature.
It is therefore important to use a transient model connected to the real-time drilling parameters in order
to more accurately estimate the downhole temperature.
However such models depend on parameters that are seldom known such as the specific heat capacity
and thermal conductivity of the various formation rocks penetrated by the well. Considering that the
specific heat capacity of sedimentary rock can easily vary between 500 J/kg.K and 1500 J/kg.K, and that
the thermal conductivity of such rocks can be found between 1.4 W/m.K to 3 W/m.K, the predicted
temperature along the borehole may deviate from reality by a few degrees (see Figure 10).
The specific heat capacity and thermal conductivity of the drilling fluid depend on the composition of
the drilling fluid. There are relatively large differences for the specific heat capacity and thermal
conductivities of different variants of the mud constituents, so if the exact thermal properties of the
individual constituents are not known then we shall account for the possible variations.
The thermo-physical properties of the drilling fluid also depend on the concentration of cuttings in
suspension (Cayeux et al. 2014a). Erroneous estimations of the cuttings concentration at various depth as
a function of time is yet another cause of discrepancies between the actual downhole temperature and the
estimated one.
Furthermore, the estimation of the amount of heat generated by the work of the bit, mechanical friction
and hydraulic turbulence is somewhat debatable. Unless there are reliable downhole temperature mea-
surements available for calibration, altogether these sources of uncertainty in the heat transfer calculations
SPE-170330-MS 11

may cause deviations in the estimation of downhole


temperature compared to actual values in the mag-
nitude of a dozen degrees Celsius.
Effect of the Drill-string Movement
Drill-string axial movement (pick-up, slack-off,
ream-up, ream-down) inside the wellbore induces
pressure pulses that propagate rapidly inside the
annulus. During such operations, the bit and drill-
string act like a piston inside the wellbore, resulting
in surge effects, which result in higher annulus
pressure, or swab effects, which result in lower
annulus pressure. Note that the acceleration and Figure 10 —Inlet and outlet temperatures when circulating at a constant
flow-rate of 2500 l/min with a typical OBM (SH stands for Specific Heat
deceleration of the string, as well as the magnitude Capacity, TC stands for Thermal Conductivity) (Courtesy Toft 2011).
of the velocity are important parameters in the gen-
eration of the pressure variations. After a transient period which duration is a function of the acceleration,
steady state conditions may be observed (if the length of a stand allow for it). With low acceleration, it
is possible to be relatively close to pseudo steady state conditions. The acceleration limit for the validity
of pseudo steady state conditions, depends more particularly on the length of the drill-string and the
compressibility of the mud. Figure 11 shows the effect of acceleration and deceleration while lifting up
a 29 meter long drill-string at 0.5m/s in a 8 ½-in hole at three different bit depths: a) 2600m MD, b) 4600
mMD and c) 7600mMD. The context in this example is a riser-less dual-gradient solution, i.e. the control
strategy is to maintain a constant interface level between the sea water and the top of the mud column in
the annulus, just above the subsea BOP Pseudo steady state conditions can be achieved with the two
lowest accelerations in the case of the shallow bit depth (2600 mMD). For the second case (bit depth ⫽
4600 mMD), only the lowest acceleration shows some characteristics of pseudo steady state conditions.
While in the case of the deepest bit depth (7600 mMD), none of the studied accelerations permit to achieve
any steady state conditions. The second observation is that the total variation of pressure at the bit depth
does not depend too much on the acceleration. Indeed, in the case a) the pressure at the bit depth shall
reduce by 3 bar simply because the bit TVD changes by 17.2 mTVD, but in fact ranges from 4 bar at the
lowest acceleration to 5 bar at the maximum studied acceleration. In the case b), the pressure at the bit
shall change by 4 bar due to the same change of bit TVD variation (17.2 mTVD), but for the variation
of pressure at the bit is 6.5 bar for the lowest studied acceleration and 8 bar for the largest calculated
acceleration. For the case c), the inclination is 81° and therefore the change of TVD is about 4.4 mTVD
which corresponds to a hydrostatic pressure difference of 1.1 bar with this particular mud, however the
pressure variation at bit during the lifting of the drill-string ranges from 6.7 bar for the lowest acceleration
to 7.5 bar for the largest acceleration. Finally, it should be noted that in the deepest bit depth case, the lift
pump flow-rate variations reach almost its limit (very close to 0) with the largest studied drill-string
acceleration, i.e. 0.03m/s2. In order to cope with larger drill-string acceleration it would be necessary to
use a larger booster pump rate.
In practice on a floater, the pseudo steady state conditions are never reached because of the
uncontrolled movement of the drill-string resulting from the movement of the vessel as a function of
heave. For calm weather conditions, the wave amplitude is about 0.5 to 1 meter with a period of 4 to 6
seconds. However in rougher sea conditions, the wave amplitude can exceed 3m in height with periods
in the magnitude of 10 to 14 seconds (see Figure 12).
As it can be seen from Figure 13, the heave movements cause larger and larger downhole pressure
variations when the amplitude of the wave increases. The example of Figure 13 shows the bottom hole
pressure 100 meter below the 9 5/8- in casing, i.e. at 7600 mMD, in the same well as described above.
12 SPE-170330-MS

Figure 11—Effects of acceleration while lifting up a 29m long stand at 0.5m/s in a 8 ½ hole at bit depth a) 2600mMD, b) 4600mMD and c) 7600mMD
in a riser-less dual-gradient setting with 1000m of water depth. A booster pump injects drilling fluid at the level of the lift pump inlet at a constant
flow-rate of 2000l/min.

Figure 12—Relationship between wave amplitude and period based on weather data from the gulf of Mexico.

In the same figure, the lift pump variations that are necessary to maintain the interface level between the
drilling fluid and the sea-water are also shown, as well as the corresponding maximum flow-rate
accelerations that the lift pump shall use to keep the level constant at the interface between the sea-water
and the drilling fluid contained in the annulus. Note that the resulting pressure variations are in most cases
larger than the one resulting from picking up a whole stand.
SPE-170330-MS 13

Figure 13—Effect of heave movements on downhole pressure in a riser-less dual-gradient solution.

Model Parameter Uncertainties


Wellbore Description Uncertainty
Any downhole pressure estimation in the open hole section of a well has to be related to a corresponding
formation layer in order to retrieve the correct geo-pressure boundaries. In practical terms, that means a
true vertical depth.
The position of a well cannot be measured directly because it is difficult to make direct spatial
measurements through the earth crust (both electro-magnetic and pressure waves are quickly attenuated
and dispersed when traveling through rock materials). In fact it is derived from the integration of
tangential measurements made along the borehole. Indeed, at a given location, it is relatively simple to
measure the inclination ␽ of the wellbore axis compared to the gravitational field and to measure the
direction (azimuth: A) of the well axis relative to the geomagnetic field or by recording direction changes
using a gyroscope. Furthermore, the length of the borehole is relatively well approximated by the length
of the drill-string, thereby providing a curvilinear abscissa s for the measurement.
The measurement of the inclination, azimuth and measured depth may be slightly biased for each
survey station. If that bias is systematic, then the errors accumulate and, after integration, result in a
potentially large discrepancy between the estimated borehole position and the true position of the well.
The importance of systematic errors in the calculation of wellbore trajectories has been reported and
analysed by Wolff and de Wardt (1981). These authors recognized six sources causing systematic errors
14 SPE-170330-MS

on the measurement of inclination, azimuth and


measured depth. The authors also developed a
method for calculating the region of space that, with
a given probability, will contain a survey station.
This region of space is defined by a 3⫻3 covariance
matrix and corresponds to an ellipsoid (see Figure
14):
(10)

Figure 14 —Ellipsoid of uncertainty at certain depth of a deviated well.

Further analysis of the problem of wellbore po-


sition uncertainty estimations has led to more ad-
vanced models. One dedicated to magnetic mea-
surements is described by Williamson (2000),
which is now considered as today’s industry stan-
dard. An extension to this model to encompass
gyroscopic measurements has been described by
Torkildsen et al. (2008). Note that the new formal-
ism still uses a covariance matrix to describe the
wellbore position uncertainty at a given survey sta-
tion. Figure 15—Quantile function for probability p of the chi-squared func-
The prediction interval for the multivariate nor- tion with 3 degrees of freedom.
mal distribution yields a region defined by:
(11)

where ⌺⫺1 represents the inverse covariance matrix, is the quantile function for probability p of
the chi-squared distribution with 3 degrees of freedom (see Figure 15), with being the mean
position of the well, and the borehole position when accounting for the uncertainty. This is the definition
of an ellipsoid in a 3-dimensional space.
So if we choose then we obtain an ellipsoid that represents the region with about a 20%
chance of containing the actual trajectory. We need to take to calculate the ellipsoid that
has a 95% chance of containing the real position of the well. Similarly, to obtain a 99% confidence of
containing the trajectory, we need to calculate the ellipsoid based on .
In that manner, for any given confidence factor, it is possible to calculate the ellipsoid of uncertainties
at each survey station of the trajectory. When investigating a certain depth of the wellbore, the algorithm
described in Appendix C allows to construct two trajectories (see Figure 16) that corresponds to the
highest and lowest position in the vertical direction for that given depth and at a given confidence factor.
It is therefore possible to analyze the influence of the wellbore position uncertainty on the downhole
pressure calculations.

Drilling Fluid Property Uncertainty


There are several types of drilling fluids. The most commonly used are the water-based mud (WBM), the
oil-based mud (OBM) and the synthetic-based mud (SBM). For these types of fluid, the density of the mud
is obtained by changing the proportion of high gravity solid (e.g. barite) in the mix. Another way of
controlling the density of the drilling fluid consists of diluting a heavy liquid such as cesium formate with
another lighter liquid such as potassium formate. The drilling fluid contains other additives in order to
manage its chemical, viscous, gelling and fluid loss control properties.
SPE-170330-MS 15

Figure 16 —Reconstruction of shallowest and deepest trajectories

Due to its compressibility and thermal expansion, the density of each of the components 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) recommend a biquadratic relationship for
the density of the components with respect to temperature and pressure (see Figure 17).
The pressure-volume-temperature (PVT) properties of the base constituents of the drilling fluid may
be quite different from one type of mud to another (see Figure 18). For instance, Zamora et al. (2013) give
a precise PVT characteristic of several base oils used in the composition of drilling fluids.
Unfortunately, there are instances where the PVT properties of the base constituent are not known. In
that case, it is necessary to consider the possible statistical variations of the PVT properties of the
constituent type, in order to evaluate its influence on the downhole pressure estimation.
A similar treatment can be made for the brine phase. However, in that case the lack of information
usually relates to the type of salts and their concentrations in the electrolyte. Kemp et al. (1989) give a
precise description on how to calculate the density of a brine containing different salts at various
concentrations. Using this model, it is possible to generate individual PVT models for the most common
solutes used in drilling fluids, i.e. NaCl, KCl and CaCl2 (see Figure 19).
If the weight fraction of the salt is not fully known or is biased by contamination from downhole
formation fluids, then the salt weight fraction shall also be statistically varied.
Similarly, it should be noted that the exact densities of the low gravity solids and high gravity solids
are usually not documented in drilling fluid or daily drilling reports.
The density of the mix is defined by a weighted average of the density of each component with the
volume fraction as a weighting factor:
(12)

where fs, fw, fo are the volume fractions of solid, brine and oil, respectively, and ␳s, ␳w, ␳o are the
respective density of solid, brine and oil. Note that the volume fractions depend also on pressure and
16 SPE-170330-MS

Figure 19 —Brine PVT for different dissolved salts for a concentration


Figure 17—Example of the density variations of a typical base oil as a of 10% in weight.
function of pressure and temperature.

Figure 18 —Example of density dependence on pressure at 50°C for


different base oils

temperature and in fact also on the density of the


components of the fluid (see Appendix D for the Figure 20 —PVT properties of an unspecified OBM with a nominal
calculation of the volume fractions). As a conse- density of 1.7 s.g. at 50°C (atmospheric pressure) having an oil/water
quence, the uncertainty on the exact density of these ratio of 70/30. The 1␴ margins correspond to a 68% confidence for the
density and the 2␴ margins represent a 95% probability of presence of
components has, in turn, a consequence on the cal- the drilling fluid density.
culation of the volume fractions and therefore on the
way the density of the drilling fluid varies as a
function of pressure and temperature.
To illustrate the effect of these uncertainties on the PVT properties of a drilling fluid, let us consider
an oil based mud with the following characteristics at 50°C and at atmospheric pressure: density 1.7 s.g.,
oil/water ratio 70/30. We do not have the PVT characteristics of the base oil and we do not know the
salinity nor the amount of low gravity solids. However, since we do know the density of the fluid at one
given temperature and pressure, we can estimate the quantities of each component by assuming some
values for the missing information:
● density of each component,
● salinity,
● fraction of low gravity solids.
By performing Monte Carlo simulations on the unknown parameters of the model, we can obtain the
most likely density variations as a function of temperature and pressure as well as their associated standard
deviations.
Assuming that the probabilistic simulations are following a Gaussian law, the variations of densities
within 1 time the standard deviation corresponds to a 68% confidence that the density is within that
SPE-170330-MS 17

Figure 21—Temperature and pressure dependence of the rheology of a water base mud measured with an Anton Paar scientific rheometer.

Figure 22—Temperature and pressure dependence of the rheology of an oil base mud measured with an Anton Paar scientific rheometer.

interval and the ones within 2 times the standard deviation are inside a confidence interval of 95% (see
Figure 20).
As for any fluid, the viscosity of drilling mud depends on the temperature but it can also depend on
the pressure as well (see Figure 21 and Figure 22). In practice, this means that the coefficients A, B and
C of the Robertson-Stiff model or ␶0, k and n of the Herschel-Bulkley model are functions of temperature
and pressure.
The rheology of the drilling fluid is measured at regular intervals (typically four times a day) and
usually this measurement is done at atmospheric pressure and at a given temperature (see Figure 23-a).
18 SPE-170330-MS

Figure 23—Estimation of the effective viscosity based on a single rheometer measurement with 3 different mud rheological models.

To retrieve the pressure and temperature dependence of the mud rheology, one relies on empirical
models for the particular drilling fluid that is in use. If the rheology dependence on pressure and
temperature of the drilling fluid is not known, or if the formulation of the mud currently used differs in
some ways from the reference model, the estimation of the effective viscosity ( ) at in situ conditions
will be uncertain. Figure 23-b, c, d shows the variation in the estimation of the effective viscosity at
various pressures and temperatures when using three different reference rheological models for the
drilling mud.
In cases where pressure and temperature dependent rheological measurements are not available, or if
there are uncertainties on the pressure and temperature dependence of the mud rheology, then it is
important to make stochastic variations of the possible rheological behavior of the mud as a function of
temperature and pressure in order to account for this source of uncertainty.
Real-time Parameter Uncertainty
The flow rate into the well is based on the pump rate multiplied by the stroke volume and pump efficiency.
However, if the motor of the mud pump is controlled as an open loop system (no encoder to return the
motor RPM), the actual motor RPM may differ from the RPM set point. Furthermore, the pump efficiency
is related to the leakage between the pistons and the cylinders. It also reflects the fluid flow back of the
inlet and outlet valves, dependent on the valve design, closing speed and condition of the valve springs.
As a consequence, the pump efficiency depends on the operating pressure and on the fluid characteristics
(density and viscosity). These considerations make it difficult to get a precise estimation of the actual flow
rate into the well.
If the flow rate in is based on the stroke-counter based pump rate instead of the motor RPM, the
measurement might be delayed by several seconds especially at low pump-rate, which often results in an
apparent SPP increase before the change of flow rate is actually measured. This makes it difficult to
compare measured and calculated values during flow accelerations.
SPE-170330-MS 19

One downhole measurement which is important for calibration of hydraulic models is the annulus
pressure measurement provided by the Pressure While Drilling (PWD) tool. Internal pressure and
temperature is typically also measured by the tool but only stored and used by the supplier. Latest
improvements in PWD technology allows pressure measurements with accuracy down to 0.1% and with
a resolution of 0.07bar. Typical operational ranges are temperatures up to 150 °C and pressure reaching
1200 bar.
Mud pulse telemetry is the most widely used way of transmitting downhole measurements to surface.
The measured values are translated into a coded signal and transmitted to surface via pressure pulses in
the drilling mud. This system requires that there is pressure in the drill-string up to the surface. Therefore,
circulation shall be established and the flow-rate shall be large enough so that the liquid level in the
drill-string reaches to the stand pipe. Mud pulse telemetry offers a relatively poor bandwidth, between
3–20 bits per second. The bandwidth dependents largely on the depth of the MWD pulser and the type
of drilling mud. Due to this constraint many of the measured variables, such as pressure, are truncated and
averaged before they are transmitted, typically with a frequency of 1–2 values per minute. However, the
transmission tool can be programmed to transmit pressure measurements with a higher frequency, at the
sacrifice of other measured variables.
During the period of time when the circulating flow-rate is too low to allow the transmission of
information, the measurements are stored and can be transmitted when circulation is resumed. Typically,
only minimum, maximum and average values of pressure are transmitted, but recent developments allows
for transmission of entire data series with a sampling interval as low as 2 seconds. This requires some
transmission time after circulation is resumed, typically 5–15 minutes to transmit pressure measurements
recorded during a 10 minutes connection. This feature opens for calibration of model parameters
associated with low circulation rates or no flow.
Mud pulse telemetry suffers from a certain delay in the signal from the downhole tool to surface. The
signal travels with the speed of sound through the drilling mud in the drill string, and is therefore
dependent on the depth of the transmission tool and the type of drilling mud. For oil base mud the speed
of sound is around 1000 –1200 m/s. This delay, in addition to the sampling interval, should be accounted
for if the PWD data is actively used to control pressure e.g. in MPD operations.
A telemetry method with bandwidth capabilities superior to the mud pulse telemetry is the wired pipe
technology. Wired pipe allows two-way communication with a bandwidth of 57.000 bits per second, and
does not use the drilling mud to transport the signal and therefore does not require circulation in order to
operate. Instead, data is transmitted through a cable embedded in the drill pipe and inductive couplers
embedded in the tool joints (Jellison et al. 2003). This allows raw data to be transmitted, and with a much
lower sampling interval than with mud pulse telemetry, typically 1 to 5. The transmission delay of that
telemetry method can be neglected; but delay can occur in the data acquisition at the rig site.
Wired drill pipe technology allows sensors and transmitters to be distributed along the drill string,
typically every 300 –350 m. This opens for more detailed measurements of the pressure in the entire
wellbore as well as mechanical forces and vibrations measurements along the drill string.

Incomplete Information
In order to perform simulations which are as accurate as possible, it is important to know how much
drilling fluid is pumped into the well at any given time. The flow-rate, which we discussed above, is of
course a crucial parameter, but we also need to know the temperature and density of the mud to be able
to calculate the mass of the drilling fluid that enters the well. The temperature can be measured in the
active pit, but there is always the risk that the pit level decreases below the sensor such that it is only the
ambient temperature that is measured. In such a situation, there is nothing wrong with the sensor, but the
information available for real-time model-based systems is nevertheless incomplete. A similar situation
20 SPE-170330-MS

Figure 24 —Change of density a) and rheology b) of a typical drilling fluid when there is a 5% concentration of cuttings in suspension.

occurs if the pit, where the sensor is mounted, is not lined up as the active pit. We will be able to read
information, but that information is not relevant.
The density in the active pit may be measured regularly, but it is usually not reported in a computer
readable format. The fluid reports that are distributed every 24 hours contain some density measurements,
but that is too late for a real-time system. There are however also potential problems with measurements
from density sensors. If they measure the density of the mud in a pit directly, we need to be sure that we
use measurements from the correct pit. If the sensors are mounted between the pit and the pump, we need
to know that the sensor we use is in the active part of the flow path.
External Factors
The drilling process produces cuttings that need to be removed from the well. With an adequate flow-rate
and drill-string rotational velocity, the cuttings are kept in suspension, even when the borehole is deviated,
hence they can be transported out of the hole together with the drilling fluid (Cayeux et al. 2014a). The
cuttings particles in suspension affect the local density and rheological behavior of the drilling fluid as a
function of their concentration and density (see Figure 24).
The determination of the cuttings concentration at a certain depth depends on what has happened to the
cuttings particle during its transport. For instance, the size of the particle changes during transport due to
the grinding action of the drill-string rotation against the wall of the borehole. Since the size of the particle
is quite important in the calculation of its slip velocity, it obviously has an impact on the cuttings
concentration at different depths. The cuttings size and shape also influences the probability that a particle
will settle on the low side of inclined sections of the well when the drilling parameters (i.e. flow-rate and
drill-string rotational speed) are not sufficient to keep it moving on the high side of the annulus.
Furthermore, the cuttings dimensions play a role in determining cuttings bed erosion. As a consequence,
there is some degree of uncertainty on the actual concentration of cuttings particles at any depth along the
wellbore (see Figure 25).
Formation rock may contain gases such as CO2, H2S or hydrocarbon gases in different proportions.
When drilling through these gas bearing formations, gas invades the drilling fluid simply because it is
released from the cuttings. Gas may dissolve in the drilling fluid, especially if the drilling fluid is an OBM
or a SBM. When dissolved, the density of the drilling fluid is slightly changed (O’Bryan et al. 1988,
Monteiro 2005). But when the pressure drops below the bubble point pressure, dissolved gases return to
gas form with bubble sizes that increase greatly as the pressure decreases, causing drastic changes in the
drilling fluid density (see Figure 26).
The concentration of the various gases is measured when the drilling fluid returns to the surface
installation, but this is too late if one wants to account for the effect of gas on the in situ drilling fluid
SPE-170330-MS 21

Figure 25—Cuttings concentration and cuttings bed height for two different cuttings particle size and cuttings bed erosion conditions.

Figure 26 —Effect of the presence of gas on drilling fluid density and rheology.

density. This leads to an additional source of uncertainty for the actual mud density at different depths of
the borehole.
Pressure loss calculations depend very much on the bulk velocity of the drilling fluid. However, in the
presence of cuttings beds or hole enlargements due to washout or hole collapse, the actual cross sectional
area of the wellbore may be quite different from the expected one. Even though a downhole caliper may
be run in some circumstances while drilling, the position of this instrument in the BHA, and therefore
close to the bit, limits the availability of the actual hole size at shallower depths. Those restrictions and
enlargements may play an important role in the pressure losses occurring through the annulus.
If mud losses occur during the drilling operation, it might not be so easy to account for this situation
in the simulations, since the location where the losses occur, and also the magnitude of the losses are not
well-known. Another problem which can be difficult to simulate is a pipe washout. Since we don’t know
the position and size of the washout, and since this type of problem modifies the flow path inside the well,
it will lead to a challenging situation when it comes to model-based systems.
Calibration
As seen in the above sections, lack of information on the characteristics of the drilling fluid (including the
side effects of contaminants such as cuttings or formation fluids), the formation rock thermal properties,
the real position of the wellbore and the actual borehole size (due to cuttings beds and hole enlargements)
are sources of uncertainty in the evaluation of downhole pressures when utilizing hydraulic models. It is
22 SPE-170330-MS

Figure 27—Variability of the downhole pressure estimations with and without downhole pressure measurements.

possible to estimate the most probable hydrostatic and circulating pressure with their associated standard
deviations by performing stochastic simulations on the various parameters that are sources of uncertainty
(see Figure 27-a). The statistical variations shall be adjusted to account for additional knowledge that may
restrain their domain of variability. The pressure estimates and standard deviations can help in determin-
ing drilling parameters that ensure a safe drilling operation.
Here is a list of typical parameters for which uncertainty on their value should be analysed with regards
to downhole pressure estimation:
● Wellbore position uncertainty (Gaussian distribution)
● Specific heat capacity of formation rock (Uniform distribution)
● Thermal conductivity of formation rock (uniform distribution)
● Specific heat capacity of base oil (Gaussian distribution)
● Thermal conductivity of base oil (Gaussian distribution)
● Specific heat capacity of brine (Gaussian distribution)
● Thermal conductivity of brine (Gaussian distribution)
● Low gravity solid specific heat capacity (Gaussian distribution)
● Low gravity solid thermal conductivity (Gaussian distribution)
● High gravity solid specific heat capacity (Gaussian distribution)
● High gravity solid thermal conductivity (Gaussian distribution)
● Cuttings specific heat capacity (Gaussian distribution)
● Cuttings thermal conductivity (Gaussian distribution)
● Factor controlling the heat generation from the bit (Gaussian distribution)
● Factor controlling the heat generation from mechanical friction (Gaussian distribution)
● Factor controlling the heat generation from turbulence (Gaussian distribution)
● Cuttings proportion (Uniform distribution)
● Cuttings density (Uniform distribution)
● Cuttings position in annulus (Uniform distribution)
● Cuttings bed position, length and height in annulus (Gaussian distribution)
SPE-170330-MS 23

● Cuttings concentration around the BHA (Uniform distribution)


● High gravity solid density (Gaussian)
● Low gravity solid density (Gaussian)
● Type of PVT for base oil (Uniform or fixed if known)
● Type of PVT for brine (Uniform or fixed if known)
● Oil/water ratio (Gaussian if known or Uniform otherwise)
● Salinity (Gaussian if known or uniform otherwise)
● Low gravity solid proportion (Gaussian if known or uniform otherwise)
● Mud pump efficiency (Gaussian)
● Surface roughness (Gaussian)
If downhole measurements are available, it is possible to further restrain the estimations. From all the
downhole pressure solutions generated using the stochastically varied parameters of the problem, we keep
only those that respect the measurement reading within its tolerance. An envelope of those depth-based
downhole pressure curves can be generated to characterize the remaining possible variations of the
downhole pressure at depths distant from the measurement. It should be noted that there exist many
possible ways to reach the downhole pressure measurement and therefore the variability of the estimated
downhole pressure may be larger at depths farther from the measurement than the one at the measurement
depth (see Figure 27-b). Two examples using that methodology, although for conventional drilling, are
described in Cayeux and Lande (2013)
If several downhole pressure measurements are made at different depths then it is possible to further
decrease the along hole uncertainty of the model predictions (see Figure 28-a). In fact along-string
measurements provide quantitative information about the possible side effects of external factors that can
influence the actual downhole pressure, like cuttings beds, various concentration of cuttings in suspension,
hole enlargements, etc. The more downhole sensors are spread along the drill-string, the more accurate
those side effects can be positioned geometrically.
In itself a downhole pressure measurement made with a certain flow allows for a reduction of the
uncertainty at other flow-rates including hydrostatic conditions, i.e. no flow. However that does not
eliminate completely the uncertainty at the measurement depth for all possible flow-rates because the
calibration point does not give enough information about what was the cause of potential discrepancies
between the model and the measurement. But if downhole pressure measurements are available at
different flow-rates for approximately the same depth and in the same drilling conditions, then it is
possible to further reduce the uncertainty of the predictions made by the model at other flow-rates (see
Figure 28-b). With mud pulse telemetry, downhole pressure measurements are not available at low
flow-rates and that has of course some consequences on the calibration of the model in the range of
flow-rate that is not covered by any measurements. We have also seen that in a dual gradient drilling
context, the main pump flow-rate has to be large enough to overcome the U-tube effect and this fact is
also setting an additional constraint to obtain downhole pressure measurement at low flow-rate. Because
wired pipe telemetry can transmit measurements regardless of any pump pressure, such technology can
provide valuable information when the main mud pumps are stopped. However, one shall bear in mind
that flow can continue for some time after the main mud pumps are stopped (because of fluid compress-
ibility or the U-tube effect) and therefore the downhole pressure measurements made in such transient
period have to be associated with predicted flow-rates.

Control Method
When a certain behavior of a physical system needs regulation such that a set of parameters (variables)
of the system follow a corresponding set of set-points, one can resort to control methods. Throughout this
24 SPE-170330-MS

Figure 28 —Increased calibration accuracy with multiple sensors and several flow-rates.

Figure 29 —A system represented by an input-output relation

Figure 30 —A control configuration for system ⌿

paper, we classify control methods into two classes, namely feed-forward and feedback. To get a clear
understanding, we need to focus on input-output relation depicted in Figure 29.
The relation between input and output can be expressed mathematically by:
(13)

where u is the input and y is the output. Here, u is the variable in the system that we can manipulate
such that we can drive y to a given set-point r as close as possible. For this control purpose, we can model
⌿ by ⌿model. One way to achieve the control goal is by setting u to be the inverse response of the model
towards the set-point as shown in Figure 30 where:
(14)
SPE-170330-MS 25

Figure 31—A feedback control configuration for system ⌿

Indeed, if we have a perfect model where ⌿model ⫽ ⌿, we achieve our control goal where y ⫽ r
because . In reality, there is no perfect model and there is always a misfit between the model
and reality. Hence, with this method, the proximity of the output y to the target r depends on the quality
of model ⌿model.
The control configuration shown in Figure 30 is an example of a feed-forward control structure. A
feed-forward control does not have an on-line mechanism to check how close the output y is with respect
to the target r. On the contrary, the other class of control method, namely feedback control, has a
correction mechanism. As shown in Figure 31, the input is manipulated by a controller K to drive (r – y)
to be as small as possible. In the extreme case, the controller K can drive y to approach its set-point r
without any model needed, provided the controller is well tuned. However, in a drilling system, variations
in the system require that the tuning is repeated on a regular basis because of the fact that a typical good
set of controller parameters only works well locally. For tuning, we need a model that captures the
physical phenomena that influence the process to be controlled. A phenomenon that is not captured in the
model used for controller tuning, is treated as uncertainty to the system. Even though a feedback controller
can accommodate with some uncertainty, there is a limit by which the control starts to be inefficient. In
a worst case scenario of the detrimental influence of uncertainty, a feedback controller may give a worse
response than a constant input. For instance, an MPD solution that uses a feedback controller in a long
well with partially compressible mud may end up with auto-sustained oscillations of the downhole
pressure, if the feedback controller is not tuned to accommodate with the delay arising from pressure wave
propagation.
A good model is indeed important for feedback design so that the uncertainty can still be accommo-
dated by the feedback. As the modelling of the drilling process for the purpose of simulation and
diagnostic starts to give better and better results, this provides the opportunity to find models that deals
with the actual complexity of modeling in real-time actual drilling operations with a relatively good
accuracy. Unfortunately, those models are not necessarily suitable for feedback control design due to their
complexity and the fact that they solve sets of partial differential equations while most of the techniques
of control design are based on ordinary differential equations. However, research on feedback control
design is very active and has made a lot of progress in the recent years.
Recent studies on backpressure MPD control strategies (Siahaan et al. 2014a and 2014b) have shown
that the advantages of feedback and feed-forward methods can be joined together by combining the two
methods in order to control the downhole pressure. In this case, the lack of high quality model to tune the
feedback control loop can be compensated by the use of a feed-forward controller using complex model.
On the opposite, the inherent lack of correction mechanism of the feed-forward method is compensated
by the use of a secondary feedback loop.
For dual gradient drilling, combination of feed-forward and feedback may be used to control the
downhole pressure. We focus on controlling the downhole pressure where the flow rate out of the lift
pump is manipulated by a controller ⌸2 such that the interface liquid level follows a target set-point. The
interface liquid level set-point is calculated by a controller ⌸1. The main focus is on the controller ⌸1
which is a combination of a feed-forward and a feedback. The controller ⌸2 that drives the lift pump is
a simple PI feedback controller.
26 SPE-170330-MS

Figure 32—Control structure for dual gradient drilling.

The feed-forward in ⌸1 is a lookup table generated by a steady state hydraulic model. The lookup table
lists the set-points of annulus level needed in order to achieve a given target of downhole pressure. The
set-point in the lookup table varies according to different values of the following variables
● main pump flow-rate (d1),
● drill string velocity (d2),
● rotary speed (d3).
Hence, the lookup table accommodates the variance in those three variables (d1, d2, d3). Variance from
other sources, for example temperature effect and cuttings, which are not accounted for in the lookup table
are considered uncertainty to the control system. As the lookup table does not capture certain uncertain-
ties, a feedback can be added as a correction mechanism to update the lookup table. There are several
means of adding a feedback mechanism to a look-up table. The standard solution is to update the steady
state hydraulic model from time to time with respect to the pressure measurement along the annulus and
to generate the lookup table from the updated steady state hydraulic model from time to time. Another way
is to update directly the lookup table. In the line of the latter method, suppose that the feed-forward in ⌸1
is given by:
(15)

where ␾lookup-table is the lookup table. We consider two ways to update directly the feed-forward in ⌸1
to manipulate the level set-point, namely
● adding a correction term to the lookup table, i.e.
(16)

where ufeedback is a feedback control to keep downhole pressure y to be as close as possible to


set-point r.
● multiplying the lookup table with a correction term, i.e.
(17)

where ufeedback is adaptively tuned by means of Whittaker’s rule (Whittaker et al. 1958).
The two methods rely on the availability of measurement of y. If the measurement is not available for
a certain period, the control scheme only utilizes the feed-forward. The performance offered by these two
control methods depend on the quality of the measurement y. Sensor noise, transmission delay, irregular
SPE-170330-MS 27

Figure 33—Reference liquid level to maintain a constant bottom hole pressure in a Low Annulus Level Return Dual-Gradient drilling solution.

sampling time, unavailability of the measurement during low circulation rate in mud pulse telemetry are
examples which can affect performance.
Figure 33 shows slices in a look-up table generated for a low annulus level return MPD solution. The
look-up table gives annulus liquid levels to maintain constant the bottom hole pressure as a function of
three parameters: the main pump flow-rate, the drill-string rotational velocity and the drill-string axial
velocity. So the look-up table is a 3-dimensional array. For the sake of simplicity, the dependence on the
drill-string axial velocity is not shown in this figure.

Conclusion
Dual-gradient drilling solutions require the management of a lift and booster pump connected to a
hydraulic network composed of several inter-related hydraulic branches. Predominantly, the flow in the
hydraulic network is in transient conditions. For long and deep wells, those transient hydraulic effects can
be both strong in magnitude and much delayed. Advanced hydraulic and heat transfer models can produce
results that are trustworthy and therefore can be used to control actively the downhole pressure in an MPD
setting. However, the accuracy of such models depends on the quality of the input data. Lack of detailed
information about the drilling fluid characteristics or the thermo-physical properties of the formation
rocks, combined with inherent well bore position uncertainty and uncertain evaluations on the cuttings
transport or the eventual presence of small quantity of gas, can be the source of substantial uncertainty in
the estimation of downhole pressures along the whole length of the open hole section. This can limit the
accuracy of the control method. Nevertheless, downhole measurements can be used to reduce these
uncertainties and therefore reach the level of accuracy that is required to perform a given drilling
operation. In fine, the need for such calibration points either distributed along the borehole or at higher
sampling interval may be a fundamental design argument of the choice of downhole measurements and
the type of downhole telemetry to be used in some deep-sea drilling operations.

Acknowledgement
The authors acknowledge the Research Council of Norway, ConocoPhillips, Det norske oljeselskap,
28 SPE-170330-MS

Statoil, Talisman, TOTAL and Wintershall for financing the work through the research center DrillWell
(Drilling and Well Centre for Improved Recovery) at IRIS.

Nomenclature
BHA -Bottom Hole Assembly
BHP -Bottom Hole Pressure
BOP -Blow Out Preventer
CFL -Courant-Friedrichs-Lewy
DGD -Dual-Gradient Drilling
DSV -Drill String Valve
LCM -Lost Circulation Material
MD -Measured Depth
MPD -Managed Pressure Drilling
OBM -Oil Base Mud
PDM -Positive Displacement Motor
PI -Proportional Integral
PVT -Pressure Volume Temperature
PWD -Pressure While Drilling
RPM -Rotation Per Minute
SBM -Synthetic Base Mud
SPP -Stand Pipe Pressure
TD -Total Depth
TVD -True Vertical Depth

Symbols
A Cross-sectional area [L2](m2)
ci Mass fraction of component i (dimensionless)
F Total mass flux [MT⫺1](kg/s)
fo Volume fraction of oil (dimensionless)
fs Volume fraction of solids (dimensionless)
fs 2
Reference volume fraction of solids at temperature T2 and pressure p2 (dimensionless)
fw Volume fraction of brine (dimensionless)
G Total mass flux per unit area [ML⫺2T⫺1](kg/m2/s)
Gi Mass flux per unit area of component i [ML⫺2T⫺1](kg/m2/s)
Gk Mass flux per unit area of phase k [ML⫺2T⫺1](kg/m2/s)
g Gravitational acceleration [LT⫺2](m/s2)
H Enthalpy per mass unit [L2T⫺2](J/kg)
Hij Element of the inverse covariance matrix [L⫺2](m⫺2)
K Friction pressure-loss term [ML⫺2T⫺2](Pa/m)
k Consistency index [ML⫺1Tn-2](Pa.sn)
m Number of junctions in the hydraulic network
ṁ Mass flow-rate per unit length source term [ML⫺1T⫺1](kg/m/s)
ṁi Mass flow-rate per unit length source term for component i [ML⫺1T⫺1](kg/m/s)
n flow behavior index (dimensionless)
p Pressure [ML⫺1T⫺2](Pa)
p1 Reference pressure for oil/water ratio [ML⫺1T⫺2](Pa)
p2 Reference pressure for volume fraction of solids [ML⫺1T⫺2](Pa)
SPE-170330-MS 29

pl Lithostatic pressure [ML⫺1T⫺2](Pa)


p0 Pressure at initial conditions [ML⫺1T⫺2](Pa)
Qc Conductive and natural convective term
Qf Forced convective term
q Source term, a mass per length per time [ML⫺1T⫺1](kg/m/s)
qs Heat generated by mechanical and hydraulic frictions [ML2T2](J)
r Set-point
Position of a point on an ellipsoid [L] (m)
Highest point in the z-direction on an ellipsoid [L] (m)
Lowest point in the z-direction on an ellipsoid [L] (m)
Centre of an ellipsoid [L] (m)
Sw Wetted perimeter of a cross-section [L](m)
s Curvilinear abscissa [L](m)
T Temperature [⌰](K)
T1 Reference temperature for oil/water ratio [⌰](K)
T2 Reference temperature for volume fraction of solids [⌰](K)
TVD True Vertical Depth [L](m)
TVD0 True vertical depth of origin [L](m)
TVDseabed Vertical depth of the seabed [L](m)
u input
t Time [T](s)
Vhgs Volume of high gravity solid [L3](m3)
Vlgs Volume of low gravity solid [L3](m3)
Average velocity [LT⫺1](m/s)
Centre of mass velocity [LT⫺1](m/s)
Centre of mass velocity of component i [LT⫺1](m/s)
Cross-sectional average velocity of a phase k [LT⫺1](m/s)
Slip velocity between phase k and l [LT⫺1](m/s)
Velocity of the inter-phase mass transfer [LT⫺1](m/s)
Velocity of the mass flux through the walls [LT⫺1](m/s)
x North coordinate [L](m)
xik Mass fraction of component i in phase k (dimensionless)
y East coordinate [L](m)
z Vertical coordinate [L](m)

Greek Letters
␣k Volume fraction of phase k (dimensionless)
Mass transfer rate per unit length between phase k and l [ML⫺1T⫺1](kg/m/s)
Shear rate [T⫺1](1/s)
␩ Effective viscosity [MT⫺1L⫺1](Pa.s)
␽ Average inclination (rd)
K1 Reference oil/water ratio at temperature T1 and pressure p1 (dimensionless)
␬ Compressibility [L⫺2T2](kg/m3/Pa)
⌳ Volume ratio of low gravity solid and high gravity solid (dimensionless)
␭ Thermal conductivity [MLT⫺3 ⌰⫺1](W/(m.K))
␮ Shear stress function [ML⫺1T⫺2](Pa)
␮kl Interface friction function [ML⫺1T⫺2](Pa)
30 SPE-170330-MS

Quantile function for probability p of the chi-squared distribution with 3 degrees of freedom
␳ Density [ML⫺3](kg/m3)
␳b Bulk density [ML⫺3](kg/m3)
␳f Cross-sectional average fluid density [ML⫺3](kg/m3)
␳i Density of component i [ML⫺3](kg/m3)
␳k Density of phase k [ML⫺3](kg/m3)
␳l Density of the liquid mix [ML⫺3](kg/m3)
␳l 2
Density of the liquid mix at reference temperature T2 and pressure p2 [ML⫺3](kg/m3)
␳hgs Density of high gravity solids [ML⫺3](kg/m3)
␳lgs Density of low gravity solids [ML⫺3](kg/m3)
␳mix Density of the mix [ML⫺3](kg/m3)
␳o Oil density [ML⫺3](kg/m3)
␳0 1
Oil density at temperature densities T1 and pressure p1 [ML⫺3](kg/m3)
␳s Combined density of low and high gravity solids [ML⫺3](kg/m3)
␳w Brine density [ML⫺3](kg/m3)
␳w 1
Brine density at temperature densities T1 and pressure p1 [ML⫺3](kg/m3)
⌺ Covariance matrix [L2](m2)
␶ Shear stress [ML⫺1T⫺2](Pa)
␶0 Yield stress [ML⫺1T⫺2](Pa)
␶w Shear stress at the wall [ML⫺1T⫺2](Pa)
␾ Body force potential [L2T⫺2](m2/s2)
⌿ Set of phase indices
⍀ Set of component indices

References
Aadnøy, B. S., 1998. Geomechanical Analysis for Deep-Water Drilling. Paper SPE 39339 presented
at the IADC/SPE Drilling Conference held in Dallas, Texas, USA, 3– 6 March 1998.
Cayeux, E., Lande, H.P., 2013. Factors Influencing the Estimation of Downhole Pressure far Away
Fom Measurment Points During Drilling Operations. Paper presented at the SIMS conference on
Simulation and Modelling, Bergen, Norway, October 16 –18, 2013.
Cayeux, E., Mesagan, T., Tanripada, S., Zidan, M., Fjelde, K.K., 2014a. Real-Time Evaluation of Hole
Cleaning Conditions Using a Transient Cuttings Transport Model. Journal paper SPE 163492 published
in SPE Drilling & Completion, vol. 29, No 1, pp. 5–21, March, 2014.
Cayeux, E., Kucs, R., & Gibson, N., 2014b. Real-Time Modelling of Drilling Using Nitrogen
Enriched Mud: A Case Study. Paper SPE 167884 presented at the SPE Intelligent Energy Conference &
Exhibition, 1–3 April, Utrecht, The Netherlands.
Corre B., Eymard R., Guenot A., 1984. Numerical Computation of Temperature Distribution in a
Wellbore While Drilling. Paper SPE 13208 presented at the SPE Annual Technical Conference and
Exhibition, Houston, Texas, USA, 16 –19 September 1984.
Courant, R., Friedrichs, K., Lewy, H. 1967. On the partial difference equations of mathematical
physics. Journal paper published in IBM Journal of Research and Development, vol. 11, No 2, pp.
215–234.
Fjelde K.K., Rommetveit R., Merlo A., Lage A., 2003. Improvements in Dynamic Modeling of
Underbalanced Drilling. Paper SPE/IADC 81636 Underbalanced Technology Conference and Exhibition,
Houston, Texas, USA, March 25–26, 2003
Frøyen, J., Savareid, O., 2000. Model equations and solution techniques for multiphase flow in pipe
networks. Research Report RF-2000/0158, International Research Institute of Stavanger, 2000.
SPE-170330-MS 31

Goldsmith, R., 1998. Mudlift Drilling System Operations. Paper OTC 8751 presented at the Offshore
Conference Technology Conference held in Houston, Texas, USA, 4 –7 May, 1998.
Isambourg, P., Anfinsen, B.T., Marken, C., 1996. Volumetric Behavior of Drilling Muds at High
Pressure and High Temperature. Paper SPE 36830 presented at the SPE European Petroleum Conference,
Milan, Italy, 22–24 Oct, 1996.
Jellison, M., Hall, D.R., Howard, D.C., Hall, H.T., Long, R.C., Chandler, R.B., Pixton, D.S., 2003.
Telemetry Drill Pipe: Enabling Technology for the Downhole Internet. Paper SPE 79885 presented at the
SPE/IADC Drilling Conference, Amsterdam, Netherlands, 19 –21 February 2003.
Kemp, N.P., Thomas, D.C., Atkinson, G., Atkinson, B., 1989. Density Modeling for Brines as a
Function of Composition, Temperature, and Pressure. Journal paper SPE 16079 published in SPE
Production Engineering, vol. 4, No 4, pp. 394 –400, Nov. 1989.
Kozicz, J. R., Juran T. L. and de Boer, L., 2006. Integrating Emerging Drilling Methods From Floating
Drilling Rigs-Enabling Drilling Solutions for the Future. Paper SPE 99135 presented at the IADC/SPE
Drilling Conference, Miami, Florida, USA, 21–23 February, 2006.
Liles, D.R., Reed, W.H., 1978. A Semi-Implicit Method for Two-Phase Fluid Dynamics. Journal
paper published in J. Computational Physics 26 1978.
Marshall D.W., Bentsen R.G., 1982. A Computer Model to Determine the Temperature Distributions
in a Wellbore. Paper 82-01-05 published in The Journal of Canadian Petroleum Technology, vol. 21, No
1, pp.63–75, Jan.-Feb. 1982.
Monteiro, E.N., 2005. Study of Methane Solubility in Organic Emulsions Applied to Drilling Fluid
Formulation and Well Control. Paper SPE 101518 presented at the SPE Annual Technical Conference and
Exhibition, Dallas, Texas, USA, 9 –12 Oct., 2005.
O’Bryan, P.L., Bourgoyne, A.T., Monger, T.G., Kopcso, D.P., 1998. An Experimental Study of Gas
Solubility in Oil-Based Drilling Fluids. Journal paper SPE 15414 published in SPE Drilling Engineering,
vol. 3, No 1, pp. 33–42, March 1988.
Ochoa, M. V., 2006. Analysis of Drilling Fluid Rheology and Tool Joint Effect to Reduce Errors in
Hydraulics Calculations. Ph.D. dissertation, Texas A&M University, Aug 2006.
Robertson R.E., Stiff H.A., 1976. An Improved Mathematical Model for Relating Shear Stress to
Shear Rate in Drilling Fluids. SPE Journal, February 1976, vol. 16, No 1, pp. 31–36.
Schubert, J. J., Juvkam-Wold, H. C. and Choe, J., 2006. Well Control Procedures for Dual Gradient
Drilling as Compared to Conventional Riser Drilling. Journal paper SPE 99029 published in SPE Drilling
& Completion, vol. 21, no. 4, pp. pp. 287–295, 2006
Siahaan, H.B., Bjørkevoll, K.S., Gravdal, J.E., 2014a. Possibilities of Using Wired Drill Pipe
Telemetry During Managed Pressure Drilling in Extended Reach Wells. Paper SPE 167856 presented at
the SPE Intelligent Energy Conference & Exhibition, Utrecht, The Netherlands, 1–3 April, 2014.
Siahaan, H.B., Vefring, E., Nikolaou, M., Gravdal, J.E., 2014b. Evaluation of Coordinated Control
During Back Pressure MPD Operations. Paper SPE 169205 presented at the SPE Bergen One Day
Seminar, Bergen, Norway, 2 April, 2014.
Stave, R., Farestveit, R., Høyland, S., Rochmann P. O. and Rolland, N. L., 2005. Demonstration and
Qualification Of A Riserless Dual Gradient System. Paper OTC 17665 at the Offshore Technology
Conference, Houston, Texas, USA, 2 May 2005.
Toft, R.E., 2011. Influence of the Thermo-Physical Parameters of Formation Rock on Temperature
Modelling, and Comparison With Recorded Data From a Drilling Operation. B. Sc. Thesis, University
of Stavanger, 2011.
Torkildsen, T., Håvardstein, S. Weston, J., Ekseth, R., 2008. Prediction of Wellbore Position Accuracy
When Surveyed With Gyroscopic Tools. Journal paper SPE 90408 published in SPE Drilling &
Completion, vol. 23, No 1, pp. 5–12, March 2008.
32 SPE-170330-MS

Trapp, J.A., Riemke, R.A., 1986. A Nearly-Implicit Hydrodynamic Numerical Scheme for Two-Phase
Flows. Journal paper published in J. Computational Physics 66 1986
Williamson, H.S., 2000. Accuracy Prediction for Directional Measurement While Drilling. Journal
paper SPE 67616 published in SPE Drilling & Completion Journal, vol. 15, No. 4, pp. 221–233, Dec.
2000.
Whittaker, H.P, Yamron, J, and Kezer, 1958. Design of Model-Reference Adaptive Control Systems
for Aircraft”, Report-164 Instrumentation Laboratory, MIT, Cambridge, MA.
Wolff, C.J.M., de Wardt, J.P., 1981. Borehole Position Uncertainty-Analysis of Measuring Methods
and Derivation of Systematic Error Model. Journal paper SPE 9223 published in Journal of Petroleum
Technology, vol. 33, No. 12, pp. 2338 –2350, Dec. 1981.
Zamora, M., Roy, S., Slater, K. S., Troncoso, J. C., 2013. Study on the Volumetric Behavior of Base
Oils, Brines, and Drilling Fluids Under Extreme Temperatures and Pressures. Journal paper SPE 160029
published in SPE Drilling & Completion Journal, vol. 28, No 3, pp. 278 –288, Dec. 2013.
Ziegler, R., Ashley, P., Malt, R., Stave, R., Toftevåg, K., 2013. Successful Application of Deepwater
Dual Gradient Drilling. Paper SPE 164561 presented at the IADC/SPE Managed Pressure Drilling and
Underbalanced Operations Conference and Exhibition held in San Antonio, Texas, USA, 17–18 April
2013.
SPE-170330-MS 33

Appendix A
Mass Conservation for Multi-component and Multi-phase flow

The mass conservation of a one-dimensional homogeneous single-phase can be written as follows (Fjelde et al. 2003):
(A-1)

where A is the cross-sectional area, ␳ is the density, is the average velocity and ṁ is the mass flow-rate per unit length
source term.
We will now consider the flow of a single component that can appear in different phase states represented by the set of
indices ⌿. Then the mass conservation for each phase of that component is expressed as:
(A-2)

where ␣k is the volume fraction of phase k, ␳k is the average density of phase k, is the average velocity of phase k, is
the mass transfer rate per unit length between phase k and l, with the convention that is the source term for phase k.
If the fluid is composed of many components appearing in different phase states, then we have the following relation:
(A-3)

where ⍀ is a set of indices representing the different components, xik is the mass fraction of component i in phase k and
ṁi is the source term of the component i.
Let us introduce the following notations:
● ␳mix is the mixture density, i.e. ␳mix ⫽ ⌺k僐⌿␣k␳k
● ci is the mass fraction of component i, ci ⫽ ⌺k僐⌿xik with ⌺i僐⍀ ci ⫽ 1
● ␳i is the density of component i, ␳i ⫽ ci␳mix. We have also ␳i ⫽ ⌺k僐⌿␣k␳k
● Gk is the mass flux per unit area of phase k:
● Gi is the mass flux per unit area of component i:
● G is the total mass flux per unit area: G ⫽ ⌺k僐⌿Gk
● is the centre of mass velocity:
● is the centre of mass velocity of the component i:
By using these notations, we can rewrite eq. (A-3) as:
(A-4)

where we used the relation


(A-5)

Equation (A-4) can be rewritten as:


(A-6)

Finally one can derive by summation over the components:


(A-7)
34 SPE-170330-MS

By introducing the mixture compressibility factor , we obtain the mass conservation equation:
(A-8)

where F is the total mass flux: .


SPE-170330-MS 35

Appendix B
Momentum Balance for Multi-component and Multi-phase flow

The Navier-Stokes equation for a one-dimensional homogeneous single-phase flow is:


(B-1)

where ␮ represents the shear stress along the walls function (which depends mostly on the fluid velocity ), ␾ is the body
force potential and is the velocity of the source term.
The generalization of the above equation to a single component that can be present in different phase states is:
(B-2)

where is the slip velocity ( ) and is the phase velocity, ␮kl is the interface friction
function for k ⫽ l and the shear stress function for k ⫽ l and is the velocity of the inter-phase mass transfer for k ⫽ l and
for k ⫽ l the velocity of the source term. The body force potential is simply the gravitational term .
Without inter-phase mass exchange velocity (i.e., ) and writing , the
momentum equation of a multi-phase flow through the pipe is the sum over the phases:
(B-3)

It should be noted that:


(B-4)

Therefore we can re-write equation (B-3) as:


(B-5)

Thus, with F ⫽ AG, and assuming a linear relation of the form , we obtain the final form of the
momentum equation:
(B-6)
36 SPE-170330-MS

Appendix C
Calculation of the Extreme Vertical Position of a Wellbore

For simplicity, the elements of the inverse covariance matrix will be denoted as Hij, @i, j 僐 [1, 2, 3]. The covariance matrix
is a symmetric matrix and therefore ⌺ ⫽ ⌺T. This property implies that the inverse matrix will also be symmetric. In practice
this gives, H12 ⫽ H21, H13 ⫽ H31 and H23 ⫽ H32. This property was utilized to derive the Cartesian equation of the ellipsoid:
(C-1)

The maximum and minimum points in the z direction are characterized by a normal to the surface of the ellipsoid that is
oriented in the z-direction and is described by:
(C-2)

where ␣ is a scalar value. This can be re-written as:


(C-3)

By solving this set of three equations, expressions for the x, y and z coordinates are given as follows:
(C-4)

where:
(C-5)

By inserting the three expressions for x, y and z in the Cartesian equation of the ellipsoid, an expression for ␣ can be
derived:
(C-6)

Due to ␣ having a positive and a negative value, two solutions and are obtained from this
set of equations. These are the points of interest, denoted for the highest point and for the lowest point.
The size and direction of an ellipsoid at a given depth of the trajectory are the results of the accumulation of the systematic
errors all along the trajectory prior to that depth. When using a single survey instrument, those systematic errors are unknown
but do not change through the integration.
We can therefore consider that the location at the surface of the ellipsoid is a characteristic of the effect of the systematic
errors on the inclination, azimuth and measured depth. By using the same relative position at each ellipsoid in the measurement
series, we can reconstruct a trajectory that bears the particular effect of those systematic errors. To find that characteristic
location, we will use the parametric coordinates of a point at the surface of the ellipsoid.
Let us consider the parametric equations of an ellipsoid in a coordinate system (X, Y, Z) that is oriented by the three axes
(a, b, c) of the ellipsoid:
(C-7)

The angles ␾ and ␪ characterize the position of a point at the surface of that ellipsoid and are specific to the systematic
errors that cause the trajectory to end up at that position of the ellipsoid. Using the same parameters ␾ and ␪ at each ellipsoid
SPE-170330-MS 37

along the wellbore, we can re-construct the trajectory that embeds the same systematic errors all along the measurements made
in that wellbore.
To obtain the coordinate system oriented by the axes of the ellipsoid, we need to diagonalize the covariance matrix and find
the transfer matrix that transforms the covariance matrix into its diagonalized version:
(C-8)

where P is the transfer matrix and D is the diagonal matrix. Note that the transfer matrix is in fact the multiplication of three
rotations necessary to transform the global Cartesian system into a local Cartesian system defined by the axes of the ellipsoid.
The following equation shows how to transform a point from the global coordinate system into the local coordinate system
attached to the axes of the ellipsoid:
(C-9)

The diagonal values of the matrix D are actually the eigenvalues of the covariance matrix. They can be found by solving
the equation:
(C-10)

where I is the identity matrix.


So to reconstruct a trajectory that corresponds to one of the extreme vertical depths associated with a particular downhole
pressure measurement, one shall apply the following procedure:
● Calculate the global coordinates of the extreme vertical point at the depth of interest using eq. C-6 and C-4
● Transform that point into the local coordinate system of the ellipsoid using the eq. C-9
● Calculate the corresponding ellipsoid parameters (␾ and ␪) using eq. C-7
● Apply the same ellipsoid parameters (␾ and ␪) at all other survey stations of the wellbore and convert the local
coordinates into the global coordinates.
38 SPE-170330-MS

Appendix D
Calculation of the Fluid Density Mix

For an OBM or an SBM, the oil/water ratio K1 is known at surface conditions (temperature T1 and pressure p1). Because the
compressibility and thermal expansion of the oil phase and the brine phase are not identical, the oil/water ratio does not remain
constant when the drilling fluid is exposed to different temperatures and pressures. At new conditions of temperature T and
pressure p, the oil/water ratio is defined by:
(D-1)

where ␳o1 and ␳w1 are the densities of the oil and brine phases (respectively) at the surface conditions, and ␳o(p, T) and ␳w(p,
T) are the densities of the oil and brine phases at the temperature T and pressure p. In view of the above discussion on the
uncertainty of the oil and brine densities, one shall also consider that the oil/water ratio is subject to uncertainty as well. Using
the above relation, we can deduce the density of the liquid emulsion:
(D-2)

where ␳l is the density of the liquid mix.


Similarly, if we have a volume ratio (⌳) between the low gravity solids (Vlgs) and the high gravity solids (Vhgs):
(D-3)

then we can express the combined solid density (␳s) as:


(D-4)

where ␳lgs is the density of the low gravity solids and ␳hgs is the density of the high gravity solids. Here, we will consider
that the density of solids is not changing with pressure and temperature even though some other investigators (Isambourg et
al. 1996) do account for those tiny variations.
The volume fraction of solids in the drilling fluid changes with the conditions of pressure and temperature because of the
compressibility and thermal expansion of the liquid phase. The estimation of the volume fraction of solid is obtained from the
initial density of the drilling fluid (␳m2) at surface conditions of temperature (T2) and pressure (p2). In those conditions the
reference solid volume fraction (fs2) is:
(D-5)

where ␳l2 is the density of the liquid in those conditions of temperature (T2) and pressure (p2). The solid volume fraction
at different conditions of the temperature and pressure is then (see Cayeux et al. 2014b for a derivation of this equation):
(D-6)

So we can calculate the density of the drilling fluid (␳mix) at any temperature T and pressure p:
(D-7)

As a consequence, the volume factions for the components are:


(D-8)

Вам также может понравиться