You are on page 1of 13

SPE-178459-MS

Management of Steam Flashing in SAGD Completion Design via the


Implementation of Flow Control Devices
Sudiptya Banerjee and Berna Hascakir, Texas A&M Univ.

Copyright 2015, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Thermal Well Integrity and Design Symposium held in Banff, Alberta, Canada, 23–25 November 2015.

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
The integrity of Steam-Assisted Gravity Drainage (SAGD) wells are often threatened by the very steam
that provides the process its name. High rate steam production, or the adiabatic flashing of steam
condensate to a vapor state, often creates an environment of ⬙flashing erosion⬙, erosive damage created by
the water-cutting effect of high velocity wet steam or the corrosive elements (e.g., carbonic acid) that can
be associated with low temperature condensate. Condensate flashing to steam also carries additional risk
from ⬙cavitation erosion⬙, erosion of completion equipment that occurs from the sudden shock wave
impacts caused by the implosion of small liquid-free zones with the condensate, similar to the well-
understood water hammer effect. Conventionally, to prevent the production of live steam or the flashing
of steam condensate, operators have relied upon choking production rates based upon monitored
temperature differences between injected and produced fluids. However, this methodology based on
subcool carries its own risks and undermines the economic viability of SAGD projects. Another approach
is to improve conformance of steam injection and fluid production by installing inflow/injection control
devices (ICDs) in either, or both, the SAGD injector and producer. If properly designed, these devices in
the producer can equalize heel-to-toe influx of bitumen emulsion, provide greater control of the sub-cool,
and behave as an autonomous (or self-regulating) valve. This presentation will focus upon this last design
objective. By coupling published performance equations for various ICD designs with steam behavior
equations, competing designs will be evaluated for their likelihood of steam flashing within a SAGD
completion and their respective ability to control flashed steam prior to the steam creating damage in the
completion. Further, the discussion will examine possible future avenues of research, such as ICD designs
with a steam throttle effect driven through a velocity-sensitive choking of steam production.
Introduction
The introduction of the Steam-Assisted Gravity Drainage (SAGD) process has revolutionized Canada’s
oil extraction industries, taking a minor player in world markets and increasing proven reserves from 7
billion to over 175 billion barrels in the last decade alone—placing Canada just behind Venezuela and
Saudi Arabia in terms of proven oil reserves (National Energy Board, 2014). These new extractable
reserves, however, are penalized by the fact that they are largely bitumen oil sands and heavy oil
reservoirs, fields that suffer a penalty in the cost of extraction, refining, and transportation, particularly
2 SPE-178459-MS

when competing with the many new shale oil plays developed in North America (Millington et al. 2014).
With these lower profit margins, operators have become understandably hesitant to introduce completion
technologies and designs that deviate from what is ⬙tried and true⬙, as the possible benefits rarely outweigh
the introduced risk to their economic model.
A notable counterexample to this hesitation is highlighted with inflow/injection control devices (ICDs)
which were introduced to the completion system of a SAGD well pair (Stalder, 2013). These completion
tools promised to correct issues with unequal steam injection profiles and to help improve the recovery
factor of the producing well during the life of a SAGD installation by adding in a variable skin effect along
the length of the horizontal well pair (Carpenter, 2014). By doing so, the new completion design enabled
uniform steam delivery, development of an ideal steam chamber, and the ability to operate at higher
production rates without steam breakthrough (Stalder, 2013). In a few short years, ICDs have gone from
unknown in Canada to having field installations in more than 60 wells and nearly 57,000 m of ICD run
in the ground.
Inflow control devices in a SAGD producer mitigates the risk of a non-uniform steam trap along the
producer well. With uniform production from all points of the lateral, a significantly lower average
subcool can be achieved while remaining certain that steam breakthrough does not occur, (because liquid
is present at all points of the lateral to serve as a mechanical diverter to vapor). By operating at lower
average sub-cool values, ICD usage also allows maximum surface area to the steam chamber, greater
interface contact, greater oil recovery rates, and minimized risk of flooding the injector well (Butler,
1994).

Design Criterias for Inflow/Injection Control Devices (ICD)


Performance Description: Restriction-Style ICD Designs
The simplest ICD design in use is the orifice ICD (Banerjee et al., 2013). These restriction-style flow
control devices introduce an obstruction in the fluid flow path that abruptly decreases the flow area,
relying on Bernoulli’s principle to provide a pressure drop (Bohra, 2004). Bernoulli equation models the
fluid flow in the orifice very well by outlining this inverse relationship between fluid velocity and fluid
pressure (Eqn A.1) (Morrison, 2003).
Orifce-style ICDs differ from variant venturi and nozzle-style ICDs due to contraction called the vena
contracta which is unique to orifice-style ICDs among all restriction-style ICDs, and will be critical to
later discussion of steam flashing in ICDs later (Darby, 2001).
In attempting to understand flow performance of restriction-style ICDs, it is helpful to try and establish
their behavior relative to better understood completion tools such as control valves. When applying
Bernoulli’s equation across a control valve, it is common to restructure the equation in terms of the valve
loss coefficient (Eqn A.2) (Darby, 2001).
In a control valve, the internal flow geometry is replicated in many modern ICDs, especially with
autonomous ICD designs (Banerjee et al., 2013) (Eqn A.3 and A.4).
In 2009, Garcia et al. published a pressure drop correlation as a function of the pressure drop
coefficient, mixture density, and mixture flow velocity, one of the first attempts to provide a performance
function that could be incorporated into a reservoir simulator (Garcia et al., 2009).
Peterson et al. (2010) built upon this work to demonstrate that an ICD pressure drop coefficient
correlates strongly to Reynolds number, allowing reservoir simulators to better determine ICD perfor-
mance for multiphase/multicomponent mixtures. In Peterson’s work, a homogenous flow model (with no
phase slippage) characterized ICD behavior using a seven-parameter function where the Reynolds number
alone was sufficient to calculate ICD pressure drop. In this work, a Reynolds dependent function will be
kept but the characteristic form for valve pressure drop is also preserved. To do so, the coefficient of
discharge, Cd, is established to be a function of Reynolds number, an intuitive deduction given the nature
SPE-178459-MS 3

of the discharge coefficient. Indeed, a literature search reveals that this relationship has already been
evaluated in the work of Reader-Harris/Gallagher (Perry, 1984) (Eqn A.5).
Thus far, the performance of restriction-style ICDs has been derived based on relationships between
four dimensionless variables: Cd, NreD, ␤, and ⌬P/P1. However, these descriptions are insufficient for
SAGD operations, as they rely on an unreasonable assumption: incompressible fluid flow (Jain et al.,
2013). In operations that frequent risk the production of live steam, methane, or even nitrogen during late-
life blow down, performance descriptions that require incompressible fluids do not adequate address the
needs of operators looking to simulate well behavior (Butler, 1994). To address this need, a fifth variable
is introduced: the expansion factor, Y, which describes the ratio of an ideal gas under adiabatic conditions
to that of the previously assumed incompressible fluid (Eqn A.6). For compressive flow through an
orifice, the expansion factor is calculated with Eqn A.7. With the help of this new term, the relationship
between flow rate and pressure can be describe as follows (Darby, 2001);
[1]

Where q is the fluid flow rate, Y is the expansion factor, Cv is the flow coefficient, A is the
cross-section area, ⌬p is the total unrecovered pressure drop across the restriction-style ICD, and ␳ is the
density of the fluid flowing through the device.
Restriction-Style ICDs and Cavitating/Flashing Liquids
For a SAGD operator, the allowable pressure drop across the ICD becomes a critical operational
restriction (Park et al., 2015). It is rare for restriction-style ICDs to be placed in series within the
completion; therefore, if the operator ⬙pulls⬙ too aggressively across a restriction-style ICD, the water
component of the produced fluid (already near vaporization temperature, as it is condensed steam) creates
a strong risk of steam flashing with no components downstream of the flashing event to control the newly
generated steam (Banerjee et al., 2013). Addressed in the introduction, producing this live steam
negatively effects the economics of SAGD production and can disrupt the operation of artificial lift
systems (Butler, 1984, Perry, 1984).
In selecting a restriction-style ICD for a SAGD completion, the lack of a vena contracta increases the
appeal of a venturi or nozzle-style ICD to mitigate risks of cavitation (and the resultant damage) within
the SAGD completion; pressures downstream of the ICD can only be lower than the outlet pressure of the
ICD if flow is continuous, so there is no opportunity for the formed gaseous ⬙bubbles⬙ to pop due to
exposure to a higher pressure as the fluid enters the production tubing. Thus, it seems reasonable to believe
that nozzle-style ICDs would be preferable to orifice-style ICDs to avoid cavitation damage and any sort
of ⬙burn through⬙ of the ICD.
In the case of live steam production, restriction-style ICDs should be optimized for the worst case
scenario: choked flow Chien, 1990). As the flow coefficient (Cv) is traditionally determined using
incompressible fluids, usually water (Banerjee et al., 2013). The calibration with air is accomplished by
introducing a new flow coefficient, Cg for gases (Eq A.8) (Darby, 2001). For varying pressures, flow rate
of steam is given:
[2]

Where ␳1 is the density of the gas at P1 with units of lbm/ft3.


Concern has been raised in the industry about how various ICD designs perform under steam
conditions; few, if any, service company has done extensive testing with steam flow through their ICD
design. More traditionally, an inert gas like air or nitrogen is used to establish gas or multiphase
performance. However, if lessons can be learned from control valve characterization, equations A.8 and
4 SPE-178459-MS

2 should be useful to extrapolate steam behavior in at least restriction-style ICD geometries when data
from steam-specific laboratory testing is absent.
Performance Description: Frictional-Style ICDs
Another common ICD type is frictional-style ICDs, devices that create their pressure drop via fluid drag
along a flow path (Banerjee et al., 2013). For this geometry, the ICD pressure drop becomes a function
of the pathway cross-sectional area and shape. Just as was done with restriction-style ICD geometries, it
becomes necessary to establish a relationship between the Reynolds number and ICD pressure drop. The
physical significance of the Reynolds number in frictional-style ICDs can be appreciated better if it is
rearranged as (Lauritzen and Martiniussen, 2011).
[3]

The Reynolds number is a ratio of the inertial momentum flux in the flow direction to the viscous
momentum flux in the transverse direction (Perry, 1984) (Equation A.9 provides the definition of Dh).
For production of bitumen emulsion, pressure drop within the frictional ICD is determined after
calculating a friction factor (Garcia, 2009). Assuming that the liquid phase behaves as a power law fluid,
the friction factor for both laminar and turbulent behavior can be captured in the equations established by
(Darby et al. (1992). Steam vapor, however, does not demonstrate power law behavior. With a viscosity
that is dependent on temperature, not on applied shear stress, steam vapor is more accurately modeled as
a Newtonian fluid (Chien, 1990, Churchill 1977, Perry, 1984). The pressure drop across the ICD is
determined as:
[4]

Where ␳ is the fluid density, V is the fluid velocity, and Dh is the hydraulic diameter. Note that many
of these frictional ICD designs will have multiple fluid channels in parallel; the preceding formula would
then be the pressure drop for a single channel, but the Reynolds number, fanning friction factor, and
velocity would have to be scaled to the proportional flow through each channel (Lauritzen and Martin-
iussen, 2011).
Frictional-Style ICDs and Cavitating/Flashing Liquids
When considering cavitation/flashing within frictional-style ICDs, an assumption is made that the steam
vapor exists under adiabatic (isentropic) conditions. This assumption is deemed to be reasonable due to
the short residence time of the fluid within the ICD.
For live steam flowing in a channel with constant cross sectional area; the mass flow rate and mass flux
must remain the same at all locations along the pipe (Eqn A.13)
As the fluid moves along the fluid channel, friction dissipates the energy of the fluid and the fluid
pressure drops just as for an incompressible fluid (Edmunds and Gittins, 1993). However, as the flow
approaches the speed of sound, a choked flow situation occurs just as previously discussed with
restriction- style ICDs (Chien, 1990).
In the case of adiabatic flow, Bernoulli’s equation for pressure loss in a pipe can be modified to
eliminate density and temperature. By using the density (Eq A.14) and temperature (Eqn A.15) expres-
sions, the mass flow rate (Eqn A.16) can be solved from Bernoulli’s equation (Eqn A.1) (Darby, 2001).
To establish the worst case scenario (choked flow), the maximum flow rate G* can be established from
the previous equation when the exit velocity hits the speed of sound, corresponding to a downstream
⬙choke pressure⬙ (P2*) as a function of the upstream pressure (P1) (Chien, 1990).
[5]
SPE-178459-MS 5

Since it is relatively straightforward to determine the pressure drop of water at saturation temperature
through a frictional-style ICD, it is a negligible exercise for the operator to determine the maximum flow
rate through the ICD without lowering outlet pressure below the vapor pressure of water, Pv. Potentially,
this choked flow behavior for frictional-style ICDs can be manipulated to maximize operator value by
providing a contingency to steam production (Park et al., 2015). The exit velocity in choked flow
conditions is a known, namely, the speed of sound. Thus, if the desired goal is to have minimal volumes
of steam enter the production tubing, the cross-sectional area of the flow channel must be decreased as
much as possible to ensure the volumetric flux is also minimized. If steam breakthrough occurs, either
through production of live steam or due to steam flashing at some part of the SAGD production lateral,
the operator may increase the drawdown in the production lateral confident in the knowledge that while
there may be additional bitumen production from the added drawdown, the volume of steam will not
increase from those zones that have already experienced steam production (Banerjee et al, 2013).
However, this same optimization has consequences, particularly during SAGD start-up. During start
up, the produced bitumen emulsion has been mobilized, but has not achieved a particularly low viscosity
(Butler, 1994). For frictional-style ICDs, the pressure drop is particularly sensitive to viscosity, which
intuitively increases the drag effects within a channel. As a result, a frictional-style ICD that has been
optimized for steam blocking will also provide an exceptionally high pressure drop across the completio
n at start-up (Banerjee et al, 2013).

Performance Description: Hybrid Autonomous Flow Control Devices


In recent years, a new class of ICDs known as autonomous ICDs have been heavily marketed as a panacea
for many operational risks in the SAGD production lateral (Carpenter, 2014). These devices demonstrate
a key behavior, the ability to increase the pressure drop and the relative choking effect for unwanted fluids
over desirable hydrocarbon. In a SAGD scenario, these devices are expected to promote the production
of bitumen-water emulsion over production of pure water and especially over live steam (Matthews,
2014). Unlike the geometries previously mentioned, there is very little homogony in design among
autonomous flow control devices. For this paper, the focus will be largely one particular design (a
hybridized autonomous ICD) that has seen significant usage in the Albertan Oil Sands (Banerjee et al.,
2013). This particular geometry has also had significant publically available data on the performance of
the ICD for a variety of different fluid. In Appendix B, performance for a 1.6FRR setting hybridized
design exposed to multiple fluids of differing viscosities is provided.
In summarizing the single and multiphase performance of this geometry, (Peterson et al. (2010)
suggested that the most effective summary equations relate a dimensionless pressure drop to Reynolds
number. To accomplish this, Peterson et al. created a seven parameter relationship to adequately describe
performance across a range of liquids and gases and their respective Reynolds number. However, this
description proves difficult to incorporate into some reservoir simulators that are designed for more
traditional control valves; thus this paper attempts to provide performance description in a similar manner
to restriction-style ICDs. Pressure drop across the ICD is provided as a function of the discharge
coefficient, with the discharge coefficient in turn a function of Reynolds number (Figure 1).
6 SPE-178459-MS

Figure 1—Relationship between Cv and Re

In attempting a polynomial regression to available single-phase data, it can be determined that the
relationship between Reynolds number and the flow coefficient may be expressed as:
[6]

Autonomous Hybridized ICDs and Cavitation/Flashing


Intuitively, the hybridized autonomous ICD should be the best geometry for managing steam flashing/
cavitation. The autonomous quality of this design increases the pressure drop for live steam production,
allowing for a more aggressive sub-cool than might be risked with more conventional completion designs.
Yet where this geometry becomes particularly interesting is in the management of steam flashing; within
the device, the fluid streamlines within the device do create a vena contracta within individual stages,
much like orifice-style ICDs do. However, as stages are stacked in series, if the pressure within the ICD
drops below the vapor pressure, Pv, the rapid transition to steam of the fluid still sees downstream chokes
and frictional pressure losses. These, combined with periods of vapor holdup and the momentary pressure
increases immediately before the restrictions of the next stage, should theoretically provide some control
of the produced steam and potential condensation of the flashed steam. However, should this not occur,
then the ICD geometry shows a choked flow behavior similar to other ICD geometries. This behavior still
needs to be confirmed with laboratory testing, as little steam performance data in ICDs has been released
to the public.
Steam Velocity Throttle - A potential design improvement?
In all ICD geometries, one detail remains consistent across all variations: if steam production through the
ICD cannot be controlled, then it will continue until a ⬙choked flow⬙ scenario occurs, where the steam
vapor is traveling through the ICD at the speed of sound (Chien, 1990).
One potential area of exploration is how to capture this significant fluid velocity to prevent steam
production (Banerjee et al., 2013). To this end, the authors suggest the exploration of venturi ejector, a
technology originally invented by Henri Giffard in 1858 as a method of feeding water to the boiler of
steam locomotives. A lifting injector uses the Venturi effect of a converging-diverging nozzle; as the
motive fluid (steam) moves through a contraction nozzle, it increases in velocity and decreases in pressure.
The localized pressure drop draws in and entrains fluid from an inlet port connected to a fluid reserve
(Darby, 2001).
If an ejector is incorporated into a SAGD-specific ICD design, then a steam flashing scenario has the
potential for the undesirable, yet high velocity vapor, to provide a lifting force that can flood the ICD
SPE-178459-MS 7

pathway with a higher density liquid (potentially drawn from the production tubing itself). In essence, the
motive energy of steam flash provides a ⬙kill switch⬙ that floods the ICD, preventing the production of live
or flashed steam through the ICD and improving the overall steam-oil ratio of the production lateral
(Perry, 1984).

Conclusions
Steam flashing/cavitation is an on-going risk regardless of ICD geometry choice. If a restriction-style ICD
is to be used, a nozzle or venturi design is preferable to a straight orifice ICD. A frictional-style ICD has
the potential to control steam flash more than many other geometries; however, optimization for steam
flash will create unreasonably high pressure drops during the high viscosity production of start-up,
suggesting this geometry is best for retrofitting a well that has been produced for some time and already
in a warm state. Hybridized autonomous geometries have potential to be an effective way to control steam
flash by ensuring downstream restriction to steam should flashing occur. This assumption, however, lacks
validation in a laboratory environment at this time. During the production of live steam, or steam from
flashing, the fluid characteristic that shows the greatest distinction between the desirable bitumen
emulsion and undesirable steam is not density or viscosity, but that of velocity. Through incorporation of
a venture ejector, the high velocity passage of steam through the ICD may provide the very mechanism
to flood the ICD pathway and ⬙kill⬙ steam production.

Acknowledgement
We acknowledge the financial support and the opportunity provided by the Texas A&M University to
conduct experiments in the Ramey Thermal Recovery Laboratory. We also acknowledge the members of
Heavy Oil, Oil shales, Oil sands, & Carbonate Analysis and Recovery Methods (HOCAM) Research
Team at Texas A&M University, Petroleum Engineering Department, for their help.

Nomenclature
A : an appropriate flow area
Cv : the flow coefficient,
Cd : the coefficient of discharge,
d1 : the smaller internal diameter, and
d2 : the larger internal diameter
Dh⫽D : (the pipe diameter) the hydraulic diameter for a fully circular pipe.
⌬Pv : the pressure drop across the valve
⌬p : the total unrecovered pressure drop across the restriction-style ICD,
G : the mass flow rate
G* : the maximum flow rate
k : isentropic exponent for steam
Kf : the loss coefficient
L1 and L2 : functions of the tap type. L1 and L2 equal to zero, if the type of the tap is corner; L1 is
one and L2 is 0.47 if the type of the tap is D and D/2; and L1 and L2 equal to 0.0254/d1,
if the type of the tap is flange.
m : the mass flux
P1 : the upstream pressure
P2 : the pressure outside the end of the channel
P2* : a downstream ⬙choke pressure⬙
Pv : the vapor pressure of water
ReD : the Reynolds number based on the larger diameter d2,
8 SPE-178459-MS

q : the fluid flow rate


V : the velocity through the area
Y : Expansion factor for nozzles
␹ : the specific heat ratio
Wp : the wetted perimeter (i.e., the length of contact between the fluid and the solid boundary)
␤ : the diameter ratio d1/d2 prior to and across the restriction,
␧ : the relative roughness of the fluid pathway
␤ : the diameter ratio d1/d2 prior to and across the restriction,

References
1. Banerjee, S., Jobling, R., Abdelfattah, T. A., & Nguyen, H. T. (2013, September 30). The Role
of Autonomous Flow Control in SAGD Well Design. Society of Petroleum Engineers. doi:
10.2118/166266-MS
2. Bohra, L. 2004. Flow and Pressure Drop of Highly Viscous Fluids in Small Orifices. M.S. Thesis,
Georgia Institute of Technology, Atlanta, Georgia. (June 2004)
3. Butler, Roger, M. (1994). Horizontal Wells for the Recovery of Oil, Gas, and Bitumen (First Ed.).
The Petroleum Society of the Canadian Institute of Mining, Metallurgy, and Petroleum. ISBN
0-96979901-2.
4. Carpenter, C. (2014, March 1). The Role of Autonomous Flow Control in SAGD Well Design.
Society of Petroleum Engineers. doi: 10.2118/0314-0136-JPT
5. Chien, S.-F. 1990. Critical Flow of Wet Steam through Chokes. J Pet Technol 42 (03): 363–370.
SPE-17575-PA. http://dx.doi.org/10.2118/17575-PA.
6. Churchill SW, Chem Eng Nov 7, 1977, p 91.
7. Darby, R., Mun, R., & Boger, DV. Chem Eng. September 1992.
8. Darby, R. Chemical Engineering Fluid Mechanics (2nd Ed.). New York, NY: Marcel Dekker,
Inc.; 2001. 304 –330.
9. Edmunds, N., & Gittins, S. D. (1993, June 1). Effective Application of Steam Assisted Gravity
Drainage of Bitumen to Long Horizontal Well Pairs. Petroleum Society of Canada. doi:
10.2118/93-06-05
10. Garcia, L., Coronado, M. P., Russell, R. D., Garcia, G. A., & Peterson, E. R. (2009, January 1).
The First Passive Inflow Control Device That Maximizes Productivity During Every Phase of a
Wells Life. International Petroleum Technology Conference. doi: 10.2523/IPTC-13863-MS
11. International Standards 2014, Measurement offluid flow by means of pressure differential devices
inserted in circular cross-section conduits running full - Part 4: Venturi Tubes, ISO 5167-4:2003,
International Organization for Standardization, Geneva, Switzerland.
12. Jain, R., Syal, S., Long, T. A., Wattenbarger, R. C., & Kosik, I. J. (2013, April 3). An Integrated
Approach To Design Completions for Horizontal Wells for Unconventional Reservoirs. Society of
Petroleum Engineers. doi: 10.2118/147120-PA
13. Lauritzen, J. E., & Martiniussen, I. B. (2011, January 1). Single and Multi-phase Flow Loop
Testing Results for Industry Standard Inflow Control Devices. Society of Petroleum Engineers.
doi: 10.2118/146347-MS
14. Matthews, C. (2014, March 1). Technology Focus: Heavy Oil (March 2014). Society of Petroleum
Engineers. doi: 10.2118/0314-0128-JPT
15. Millington, D., Murillo, C., and McWhinney, R. ⬙Canadian Oil Sands Supply Costs and Devel-
opment Projects (2014-2048). Canadian Energy Research Institute. July 2014.
16. Morrison, G.L. 2003. Euler Number Based Orifice Discharge Coefficient Relationship. J. Fluids
Eng. 125 (1): 189 –191. http://dx.doi.org/10.1115/m521955.
SPE-178459-MS 9

17. National Energy Board. 2014. Canadian Energy Dynamics: Review of 2014. February 2015,
http://www.neb-one.gc.ca/nrg/ntgrtd/mrkt/dnmc/2014/2014nrgdnmc-eng.pdf (accessed 11 June
2015.
18. Park, C., Choi, J., Lee, C., Ahn, T., & Jang, I. (2015, July 27). Operation Constraints of Steam
Assisted Gravity Drainage Considering Steam Interference to Accomplish Optimum Bitumen
Recovery. International Society of Offshore and Polar Engineers.
19. Perry, Robert, H. and Green, Don, W. (1984). Perry’s Chemical Engineers’ Handbook (Sixth
Ed.). McGraw Hill. ISBN 0-07-049479-7.
20. Peterson, E. R., Coronado, M. P., Garcia, L., & Garcia, G. A. (2010, January 1). Well Completion
Applications for the Latest-Generation Low-Viscosity-Sensitive Passive Inflow-Control Device.
Society of Petroleum Engineers. doi: 10.2118/128481-MS.
21. Stalder, J.L. ⬙Test of SAGD Flow-Distribution-Control Liner System in the Surmont Field,
Alberta Canada⬙. JCPT, March 2013, Vol. 52, No. 2.
10 SPE-178459-MS

Appendix A

[A.1]

[A.2]

[A.3]

[A.4]

[A.5]

[A.6]

[A.7]

[A.8]

C1⫽Cg/Cv and is determined by measurements on air.


[A.9]

[A.10]

[A.11]

[A.12]

[A.13]

[A.14]

[A.15]
SPE-178459-MS 11

[A.16]
12 SPE-178459-MS

Appendix B

Table B.1—FRR Hybridized Geometry Performance


⌬P Oil Flow
[psi] [gpm]

Viscosity Standard Standard


Fluid [cP] Specifc Gravity Average Deviation Average Deviation

LVT-200 2.5 0.82 14.84 0.1 3.31 0.01


LVT-200 2.5 0.82 29.46 0.1 4.68 0.01
LVT-200 2.5 0.82 44.29 0.09 5.74 0.01
LVT-200 2.5 0.82 72.7 0.14 7.37 0.01
LVT-200 2.5 0.82 87.66 0.13 8.09 0.01
LVT-200 2.5 0.82 146.26 0.45 10.5 0.01
Water 1 1 15.19 0.4 3.06 0.04
Water 1 1 29.33 0.06 4.26 0.01
Water 1 1 43.99 0.09 5.21 0.01
Water 1 1 72.6 0.21 6.71 0.01
Water 1 1 87.05 0.48 7.35 0.02
Water 1 1 146.55 0.21 9.58 0.01
Ultra-S2 10 0.825 14.53 0.17 3.34 0.03
Ultra-S2 10 0.825 28.63 0.09 4.66 0.01
Ultra-S2 10 0.825 43.57 0.13 5.73 0.01
Ultra-S2 10 0.825 72.37 0.19 7.36 0.01
Ultra-S2 10 0.825 87.42 0.23 8.08 0.01
Ultra-S2 10 0.825 144.24 0.52 10.38 0.01
80N 20 0.853 14.49 0.13 3.33 0.01
80N 20 0.853 29.62 0.18 4.75 0.02
80N 20 0.853 43.46 0.36 5.72 0.03
80N 20 0.853 73.15 0.16 7.37 0.01
80N 20 0.853 87.47 0.27 8.03 0.02
80N 20 0.853 146.55 0.96 10.39 0.04
600N 200 0.8765 14.83 0.05 3.33 0.01
600N 200 0.8765 28.94 0.49 4.78 0.05
600N 200 0.8765 44.32 0.36 5.92 0.01
600N 200 0.8765 73.21 0.42 7.57 0.02
600N 200 0.8765 87.3 0.29 8.26 0.01
600N 200 0.8765 145.46 0.64 10.57 0.03
SPE-178459-MS 13

Table B.2—Calculated Reynolds Number and Flow Coefficient


Fluid ⌬P [psi] Oil Flow [gpm] Cv Re

LVT-200 14.84 3.31 0.7781 9394.058


LVT-200 29.46 4.68 0.7808 13282.23
LVT-200 44.29 5.74 0.781 16290.6
LVT-200 72.7 7.37 0.7827 20916.68
LVT-200 87.66 8.09 0.7824 22960.1
LVT-200 146.26 10.5 0.7862 29799.88
Water 15.19 3.06 0.7851 26477.25
Water 29.33 4.26 0.7866 36860.48
Water 43.99 5.21 0.7855 45080.54
Water 72.6 6.71 0.7875 58059.58
Water 87.05 7.35 0.7878 63597.31
Water 146.55 9.58 0.7914 82892.82
Ultra-S2 14.53 3.34 0.7959 2384.25
Ultra-S2 28.63 4.66 0.791 3326.529
Ultra-S2 43.57 5.73 0.7885 4090.345
Ultra-S2 72.37 7.36 0.7858 5253.916
Ultra-S2 87.42 8.08 0.7849 5767.886
Ultra-S2 144.24 10.38 0.785 7409.735
80N 14.49 3.33 0.808 1228.895
80N 29.62 4.75 0.8061 1752.928
80N 43.46 5.72 0.8014 2110.894
80N 73.15 7.37 0.7959 2719.806
80N 87.47 8.03 0.793 2963.371
80N 146.55 10.39 0.7927 3834.299
600N 14.83 3.33 0.8096 126.275
600N 28.94 4.78 0.8319 181.2597
600N 44.32 5.92 0.8325 224.489
600N 73.21 7.57 0.8283 287.0577
600N 87.3 8.26 0.8277 313.2228
600N 145.46 10.57 0.8205 400.819

You might also like