Você está na página 1de 10

OTC 19514

Predicting When and Where Hydrate Plugs Form in Oil-Dominated


Flowlines
J ohn Boxall, Simon Davies, Carolyn Koh, and E. Dendy Sloan. Center for Hydrate Research, Colorado School of
Mines
Copyright 2008, Offshore Technology Conference

This paper was prepared for presentation at the 2008 Offshore Technology Conference held in Houston, Texas, U.S.A., 58 May 2008.

This paper was selected for presentation by an OTC 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 Offshore Technology Conference and are subject to correction by the author(s). The material does not necessarily reflect any position of the Offshore Technology Conference, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Offshore Technology Conference 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 OTC copyright.


Abstract
This work provides a means to predict when and where hydrate plugs will form in oil-dominated flowlines. The method was
funded by DeepStar and is based on a hydrate kinetic model, CSMHyK, developed over the last six years, which is currently
an addition to the transient multiphase program OLGA by SPT (Scandpower) Inc. The predictions show good agreement to
data for hydrate formation in three flowloops with five oils.

Recent CSMHyK-OLGA workshops have been held in Houston and Oslo, and major companies are beginning to use the
program in flow assurance to predict where and when hydrate plugs will form in flowlines.


Introduction
Gas and oil sub sea production and transportation are moving to deeper developments where the temperature and pressure
conditions are well within the hydrate stability region. The subsequent increased risk of hydrate formation requires new
strategies in flow assurance. Traditional methods of thermodynamic avoidance are impractical or uneconomic due to the
large amounts of thermodynamic inhibitor (e.g. methanol or monoethylene glycol) required to prevent hydrates from forming
under these conditions [1, 2]. Transient operations are particularly problematic due to the temporarily extreme subcoolings
under these conditions. The prediction of hydrate formation rates in these conditions is a major challenge requiring
knowledge of the kinetics of hydrate formation, rather than only hydrate thermodynamics. The ability to predict the rate of
hydrate formation in sub sea flowlines under restart and shutdown conditions is extremely valuable in establishing new
operating procedures during transient operations and in flowline design.

CSMHyK is a subroutine module for the OLGA (SPT Group) multiphase flow simulator. Researchers at CSM in
cooperation with the SPT Group have been developing the module since 2003. The model predicts the rate of hydrate
formation using a first-order rate equation based on the thermal driving force. The rate equation was originally proposed by
Vysniauskas and Bishnoi [3] in the absence of mass and heat transfer limitations. In order to accurately simulate hydrate
formation in flowloops, it was necessary to reduce the intrinsic kinetic constant by a factor of 500 [4]. The adjusted
parameter accounts for mass and heat transfer limitations in the flowloops. The current model assumes that the hydrate
particles convert directly from emulsified water droplets. Nucleation is assumed to occur instantaneously at a subcooling of
6.5F, a parameter proposed by Matthews [5]. Once formed, the model assumes that these particles remain in the oil phase.
The change in relative viscosity of this phase is then found from the Camargo and Palermo [6] correlation for steady state
slurry flow. An overview of the current CSMHyK module and its integration into OLGA is shown in Figure 1.

2 OTC 19514
Driving Forces
Yes
No
Dissociation
Model
Pipeline
Section n-1
Pipeline
Section n+1
Amount of Hydrate
Relati ve Viscosi ty
Growth Rate
Formation
Model
Thermo Properties
System Properti es
Thermo Properties
Hydrate
forming?
CSMHyK CSMHyK
Output
Flui d Properties
Pipeline
Section n
OLGA


Figure 1: Integration of CSMHyK into OLGA

The CSMHyK plugging model has the following characteristics:
1) Water is assumed to be dispersed in the oil phase.
2) Hydrate growth rate depends on the interfacial area between water and oil. Surface area is calculated from OLGA.
3) The size of the water droplets is fixed to one single value which can be modified by the user (default 40 m).
4) The primary hydrate particles agglomerate in the oil phase. The size is determined from the Camargo and Palermo
force balance [6]

of adhesive and shear forces. In practice this gives a single agglomerate size for a given hydrate
particle concentration in the slurry at a given shear rate.
5) Plugging is caused by the viscosity increase of the hydrate slurry from increasing hydrate volume fractions.
Plugging is defined by the attainment of a preset maximum allowable viscosity value.
6) The model does not account for hydrate deposition on the wall, and all hydrate growth occurs in the liquid phase.
Deposition can, however, be simulated by imposing a low slip value between the hydrate and oil. In this case the
hydrate particles will have a low velocity and accumulate where they initially form.
7) The model accounts for the exothermic heat generated from hydrate formation.

The key challenge was to devise a strategy for testing the kinetic model, CSMHyK-OLGA. In the absence of oil dominated
flowline data. The model was successfully tested against experimental flowloop data in three different flowloop facilities
(Texaco, ExxonMobil (XoM), and the University of Tulsa [4, 7]).


Model Verification by Comparison with Flow Loop Experiments using CSMHyK-OLGA
The updated CSMHyK model has been shown to successfully predict hydrate formation in the ExxonMobil flow loop with
the intrinsic kinetic rate decreased by a factor of 500. This fitted rate constant was then used to correctly predict hydrate
formation in the University of Tulsa flow loop for additional oils without adjustment. The value of the adjustable parameter
depends on the way the surface area for hydrate formation is calculated. In this case, the surface area was calculated based on
the droplet surface area (log-normal distribution with a mean size of 40 microns) assuming full dispersion of the water phase
in the oil. The reduced formation rate constant (0.002 times the laboratory-measured rate constant) suggests that hydrate
formation (in the flow loop) is not limited by intrinsic kinetics, but instead that mass transfer or heat transfer, limitations
exist.

Figure 2 shows examples of the comparison between the CSMHyK-OLGA model and the flow loop data for experiments in
the XoM flow loop (a) and the University of Tulsa flow loop (b, c, d). The fit for the XoM flow loop (Figure 2a) is an
example of one out of 22 experiments with a single oil (Conroe crude) varying water cut, pump speed and liquid loading.
The University of Tulsa examples (Figure 2b, c, d) are examples from approximately 12 experiments with four different oils.
Further information on the CSMHyK-OLGA simulations of flow loop experiments can be found in references [4, 7]

OTC 19514 3

Figure 2. Graphs Showing Rate Constant Fit to XoM Flow Loop (a) Predicted Formation in TU Flow Loop (b, c, d).


Industrial Use of CSMHyK-OLGA
CSMHyK-OLGA has been used by Chevron to model a typical flow line scenario, and was shown to be useful to flow
assurance design in its elementary form. Design parameters such as insulation (overall heat transfer coefficient) and
operational changes such as water cut were investigated. A startling prediction result was the self-limiting ability of hydrate
formation, due to the exothermic hydrate formation process. Even with minimal insulation, the flow line temperature was
predicted to increased to the hydrate equilibrium point after very little hydrate formation. Further formation is then limited by
removal of heat from the system.

Base Case Descriptions
To explore the usefulness of CSMHyK-OLGA, a simplified tieback and riser (Figure 3) was modeled by Chevron [7], with
conditions similar to those found in the Gulf of Mexico. A thirty mile horizontal tieback is connected to a vertical riser, 5000
ft in length, which in turn is connected to a separator. The well head was assumed to be at 150 F and 4000 psig, with the sea
floor temperature at 39.2 F. At the surface, the ambient temperature was 72 F, and the pressure at the separator was set to
250 psig. The pipe had an 8 diameter and an overall heat transfer coefficient of 4 BTU/hr ft
2
F. The oil flowing through
the pipe was 38 API Gulf of Mexico, the water cut was 30%, and the gas:oil ratio was 1000.
P
FWH
=4000 psi
T =150 F
P
separator
=250 psig
30 miles T =39.2 F
T =72 F
P
FWH
=4000 psi
T =150 F
P
separator
=250 psig
30 miles T =39.2 F
T =72 F

Figure 3. Schematic of the Tieback.

An additional base case was used to examine the effect of water cut on the system (Figure 4). A section of the subsea
pipeline was simulated, starting near hydrate formation conditions with the overall heat transfer coefficient increased to 20
BTU/hr ft
2
F. The oil, gas: oil ratio, subsea temperature, and pipe diameter were kept the same, but the water cut was varied
between 1%, 10%, and 30%. A 10,000 m section was simulated where the entrance was held at 2000 psig and 80 F and the
end section was held at 1500 psig.

(a) (b)
(c) (d)
Conroe 550rpm 35% watercut
CSMHyK-OLGA
Experiment
5000 ft
4 OTC 19514
10,000 m
P = 2000 psig
T = 80 F
P = 1500 psig

Figure 4. Schematic of the Subsea Pipeline Section.

Results and Conclusions
The following results were found by running the simulation on the tieback and riser. It was found that not all the water/gas
was completely converted to hydrate. Instead, only a few percent volume fraction of hydrate conversion occurred. The
conversion was sufficient to cause the temperature inside the pipeline to increase back to the hydrate formation temperature
due to the exothermic hydrate formation process (Figure 5). Further formation was then limited by the ability of the system
to remove heat, and the hydrate formation proceeded at a rate sufficient to keep the temperature at the hydrate equilibrium
temperature. This suggests that hydrate formation in this subsea pipeline example is heat transfer limited.


Figure 5. Temperature Profile Plot of the Tieback.

In the subsea pipeline simulations, the heat transfer limitation was again confirmed, with the pipeline temperature increasing
to the hydrate formation temperature upon hydrate formation. In addition, it was found that this hydrate formation
temperature would be maintained until all the water in the pipe was converted into hydrate (Figure 6).
30
40
50
60
70
80
90
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Distance (m)
T
e
m
p
e
r
a
t
u
r
e
(F
)
ambient T
T
hydrate
formation T
1 %
30
40
50
60
70
80
90
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Distance (m)
T
e
m
p
e
r
a
tu
r
e
(F
)
ambient T
T
hydrate
formation T
10 %
30
40
50
60
70
80
90
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Distance (m)
T
e
m
p
e
ra
tu
re
(F
)
ambient T
T
hydrate
formation T
30 %

Figure 6. Subsea Pipeline Temperature Profile (1%, 10% and 30% Water Cut).

These simulations demonstrated that hydrate formation in a subsea tieback with 30% water cut could be heat transfer limited.
In addition, the results show that the water cut can be a controlling factor in determining how much the cooling of a pipeline
section can be buffered y hydrate formation.


Application Example 1: Determining the Hydrate Blockage Potential for a Deepwater Tieback
In this example, a fictional subsea tieback is simulated, and the effect of adjusting four key parameters is investigated. The
fictitious deepwater tie-back for the simulation is a 30 mile long pipeline in 5000 ft of water. The flow line barely enters the
hydrate domain. Figure 7 illustrates some of the basic inputs for the OLGA model generated in this example.
P
a
3e7
2.5e7
2e7
1.5e7
1e7
F
150
140
130
120
110
100
90
80
70
60
50
Length [m]
50000 45000 40000 35000 30000 25000 20000 15000 10000 5000 0
P
T
tieback
Hydrate
formation T
Length along Tieback [m]
P
r
e
s
s
u
r
e

[
P
a
]

T
e
m
p
e
r
a
t
u
r
e

[
F
]

OTC 19514 5
Wellhead
Separator
Riser
Flow Line
5000 ft
30 miles, 10diameter
P =15,000 psia
T =220 F
Water cut =35%
T =39.2 F
U =4 BTU/Ft
2
F Hr
P =300 psia
T =80 F
Well
Area of Hydrate Risk Area of Hydrate Risk

Figure 7: Schematic of the Tieback in the Simulation


Adjustable Parameter Description
This example demonstrates how the adjustable parameters can be used to evaluate the blockage potential of a fictional
tieback. The four main adjustable parameters in the CSMHyK-OLGA model are as follows {OLGA Keyword in block
capitals and applicable adjustment range in square brackets}:
1. K1SCALINGFACTOR: Adjustment of the intrinsic kinetic formation constant [0-1, default = 1 (formation rate =
laboratory-measured rate constant)]
2. SUBCOOLING: Adjustment of the sub-cooling temperature (hydrate equilibrium T system T) before hydrate
formation initiates [0+, default = 6.5F]
3. SOIL: Adjustment of the velocity difference (slip) between the oil and the hydrate [0-1, default = 1 (V
oil
=V
hyd
)]
4. SIZESCALING: Adjustment of the individual hydrate (monomer) particle size used in the aggregation model [0+,
default = 1, hydrate particle size = 40 m]

The keyword K1SCALINGFACTOR refers to a multiplier for the intrinsic kinetic rate. A value of 0.01 refers to slowing the
intrinsic kinetic rate by a factor of 100.

The keyword SUBCOOLING is the extent that the system temperature must fall below the hydrate equilibrium temperature
for hydrates to start forming (i.e. SUBCOOLING = T
eqm
T
system
). Once hydrates are present, they can continue to form as
long as the system temperature is below the hydrate equilibrium temperature. This meta-stability is almost always seen in
hydrate nucleation [1]. The default value for the SUBCOOLING is 6.5F, which is taken from Matthews et al.

[5].

The keyword SOIL is the term given to the slip factor between the oil velocity and the hydrate velocity. Alternatively,
SWAT can be used which adjusts the slip between the water velocity and the hydrate velocity. The default is 1 for SOIL and
0 for SWAT. The calculation of the hydrate velocity is based on the following equation, where V refers to the superficial
velocity of the phase (in the subscript):

The keyword SIZESCALING refers to the multiplication factor of the monomer hydrate particle size; the default value is 40
m.

Simulation Results
Figure 8 shows the simulated temperature profile and hydrate formed using the default values, i.e. 6.5F sub-cooling, intrinsic
kinetics, oil slip factor of 1 (no deposition), and 40 m droplet/hydrate particle size. Hydrate formation in this case is heat
transfer limited, i.e. limited by the ability of the pipeline to loose heat to the surroundings due to the exothermic heat of
hydrate formation. This is seen by the pipeline temperature quickly rising to the hydrate equilibrium temperature. Figures 9
to 12 show the effect of changing each of the parameters, the most sensitive of which is the oil slip.

6 OTC 19514

Pipeline T
Equilibrium T
Hydrate fraction

Figure 8: Simulation Results using Base Case Values


Pipeline T
Equilibrium T
K1SCALING = 1.0
K1SCALING = 0.1
Pipeline T K1SCALING = 1.0
K1SCALING = 0.01
Equilibrium T

Figure 9: The K1 Scaling Factor must be reduced by 100 to Move the System from Heat Transfer Control (solid line
represents the default case)

SUBCOOLING = 0.5 F
SUBCOOLING = 4 F
SUBCOOLING = 6.5 F
SUBCOOLING = 9 F
SUBCOOLING = 0.5 F
SUBCOOLING = 4 F
SUBCOOLING = 6.5 F
SUBCOOLING = 9 F

Figure 10: Changing the Subcooling Changes the Point of Onset of Hydrate Formation, but has Little Effect on the Amount
of Hydrate Formed


Figure 11: The Oil Slip is the Most Sensitive Parameter; A Low Value for SOIL Leads to Hydrate Accumulation (solid line
represents the default case)
SOIL = 0.01
plugged
SOIL = 0.01
plugged
SOIL = 0.01
plugged
SOIL = 0.01
plugged
OTC 19514 7
SIZESCALING = 2
SIZESCALING = 0.5
SIZESCALING = 2
SIZESCALING = 0.5

Figure 12: The Droplet Size has Little Effect on the Amount of Hydrate Formed or on the Pressure Drop; Increasing the
Droplet Size Reduces the Relative Viscosity (solid line represents the default case)

Application of the CSMHyK-OLGA model to a typical tie-back model resulted in the following conclusions, with regards to
the sensitivity of the model to the four main adjustable parameters.
1. Hydrate / oil velocity slip factor (keyword: SOIL) has the greatest effect on the plugging tendency of a pipeline.
2. Hydrate formation in an oil dominated pipeline system is generally heat transfer limited (rather than limited by
intrinsic kinetics)
Kinetic rate constant must be reduced by more than 100 times for the system to not be limited by heat
transfer through the pipe wall. This is shown by the system temperature being equal to the hydrate
equilibrium temperature with smaller adjustments to the kinetic rate equation.
3. Sub-cooling before initial hydrate nucleation has little effect on the overall hydrate formation
Only has an effect on the onset location
4. Hydrate particle monomer size has little effect on hydrate formation or plugging


Application Example 2: Assessing Proposed Procedures for the Cold Restart of an Oil Well
The second example problem demonstrated in the CSMHyK-OLGA workshops involved the simulation of the restart of a
vertical well which had been shut-in for five days. A schematic representation of the well is shown in Figure 13. The water
depth was 1700 ft and the base of the well was located 5000 ft below the mud line. The reservoir conditions in the simulation
were 240F and 5000 psig. The simulation was performed in 4 parts:
Part A: Establish steady state operation for 5 days
Part B: Shut-in for 5 days
Part C: Add the CSMHyK module to the OLGA model
Restart the well
Assess the plugging potential
Part D: Repeat with methanol injection

Parts A and B are critical to ensure that a representative fluid distribution and temperature profile are obtained prior to restart.

1700 ft
Choke
Separator
Well
SSSV
Wellhead
5000 ft
240 F
5000 psi
Area of Hydrate Risk Area of Hydrate Risk

Figure 13 A Schematic Representation of the Well to be Simulated

Results from the Steady State and Shut-in Simulations
The simulated temperature-pressure profile of the well at the end of the five day shut-in is shown in Figure 14. The
hydrate equilibrium curves for various concentrations of methanol are plotted on the same figure. Figure 14 shows that even
30 wt% of methanol is insufficient to fully thermodynamically inhibit the well, since hydrates are still stable at the pressure
and temperature conditions in the well at the mud line. However, it is possible that the line will be kinetically inhibited; i.e.
8 OTC 19514
there will be insufficient time and / or subcooling to form enough hydrate to plug the well. This is the main advantage of a
time-dependent hydrate formation model incorporated into a transient multiphase flow simulator compared to steady-state
thermodynamic models.

0
500
1000
1500
2000
2500
3000
3500
4000
30 40 50 60 70 80
Temperature (F)
P
r
e
s
s
u
r
e

(
p
s
i
)5 wt %
10 wt %
15 wt %
20 wt %
25 wt %
30 wt %
No MeOH
Shutin/Restart
Reservoir
Mudline
Wellhead

Figure 14 A Pressure-Temperature Plot Showing the Hydrate Equilibrium Line for Various Methanol Concentrations and
the Well Pressure-Temperature Profile Following the Shut-in

Restart Results for Various Methanol Concentrations
The CSMHyK model was incorporated into the restart model. An oil slip ratio of zero was used, which forced hydrate
particles to accumulate where they initially formed rather than being transported out of the well by the oil phase. A
subcooling criterion for nucleation of 6.5
o
F was used, based on Matthews et.al. [5] observations. Intrinsic growth kinetics
were used for the simulations, i.e. no mass or heat transfer limitations.

The results from the restart simulations are shown in Figure 15 for methanol concentrations of 0, 10, 20 and 30 wt%. In
each case, the pressure-temperature profile started off partially inside the hydrate stability region. In the uninhibited case, the
accumulation of hydrate in the well caused a significant reduction in the oil and gas flow rates. This meant that warm
reservoir fluids took longer to reach the hydrate and dissociate it. After 30 minutes the well was still deep inside the hydrate
formation region. With 30 wt% of methanol, the well was less than 5 minutes in the hydrate formation region, and the
maximum subcooling was just 5
o
F which did not satisfy the nucleation criterion.

0
500
1000
1500
2000
2500
3000
3500
4000
30 40 50 60 70 80
Temperature (
o
F)
P
r
e
s
s
u
r
e

(
p
s
i
)
Equilibrium Curve for No MeOH
0 minutes
5 minutes
10 minutes
15 minutes
30 minutes
0
500
1000
1500
2000
2500
3000
3500
4000
30 40 50 60 70 80
Temperature (
o
F)
P
r
e
s
s
u
r
e

(
p
s
i
)
Equilibrium Curve for 10 wt% MeOH
0 minutes
5 minutes
10 minutes
15 minutes
30 minutes

0
500
1000
1500
2000
2500
3000
3500
4000
30 40 50 60 70 80
Temperature (
o
F)
P
r
e
s
s
u
r
e

(
p
s
i
)
Equilibrium Curve for 20 wt% MeOH
0 minutes
5 minutes
10 minutes
0
500
1000
1500
2000
2500
3000
3500
4000
30 40 50 60 70 80
Temperature (
o
F)
P
r
e
s
s
u
r
e

(
p
s
i
)
Equilibrium Curve for 30 wt% MeOH
0 minutes
5 minutes 10 minutes

Figure 15 Pressure-Temperature Plots Showing the Hydrate Equilibrium Line and the Well Conditions at Various Times
after Restart for 0, 10, 20 and 30 Wt% Methanol

OTC 19514 9

Profile plots for the well in the four simulations are shown in Figures16 to 19. The most severe flow impedance occurred in
the uninhibited case, as indicated by the low flow rate of the gas phase. No hydrate formed in the simulation with 30 wt% of
methanol.

Figure 16 A Profile Plot from the CSMHyK-OLGA Simulation Showing a Hydrate Plug Forming in the Uninhibited Well
(Indicated by a Viscosity Spike and a Drop in Flow Rate)


Figure 17 A Profile Plot from the CSMHyK-OLGA Simulation Showing a Hydrate Plug Forming in the Well with 10 Wt%
Methanol (Indicated by a Viscosity Spike and a Drop in Flow Rate)


Figure 18 A Profile Plot from the CSMHyK-OLGA Simulation Showing a Hydrate Plug Forming in the Well with 20 Wt%
Methanol (Indicated by a Viscosity Spike and a Drop in Flow Rate)

0
500
1000
1500
2000
2500
3000
3500
4000
4500
5000
0 1000 2000 3000 4000 5000 6000 7000
Distance from Reservoir (ft)
P
r
e
s
s
u
r
e

(
p
s
i
a
)

O
R

V
i
s
c
o
s
i
t
y

R
a
t
i
o
0
2
4
6
8
10
12
14
16
V
e
l
o
c
i
t
y

(
f
t
/
s
)
Pressure
Viscosity Ratio
Gas Velocity
Liquid Velocity
0
500
1000
1500
2000
2500
3000
3500
4000
4500
5000
0 1000 2000 3000 4000 5000 6000 7000
Distance from Reservoir (ft)
P
r
e
s
s
u
r
e

(
p
s
i
a
)

O
R

V
i
s
c
o
s
i
t
y

R
a
t
i
o
0
2
4
6
8
10
12
14
16
V
e
l
o
c
i
t
y

(
f
t
/
s
)
Pressure
Viscosity Ratio
Gas Velocity
Liquid Velocity
0
500
1000
1500
2000
2500
3000
3500
4000
4500
5000
0 1000 2000 3000 4000 5000 6000 7000
Distance from Reservoir (ft)
P
r
e
s
s
u
r
e

(
p
s
i
a
)

O
R

V
i
s
c
o
s
i
t
y

R
a
t
i
o
-2
0
2
4
6
8
10
12
14
16
V
e
l
o
c
i
t
y

(
f
t
/
s
)
Pressure
Viscosity Ratio
Gas Velocity
Liquid Velocity
10 OTC 19514

Figure 19 A Profile Plot from the CSMHyK-OLGA Simulation Showing no Hydrate Plug Forms in the Well with 30 Wt%
Methanol

The most sensitive parameter in the simulation is shown to be the oil slip ratio; a slip ratio of zero represents the worst case
and is recommended for most industrial applications. With a slip factor of zero, CSMHyK-OLGA can predict hydrate plug
formation in industrial systems; a plug is indicated by a spike in the viscosity ratio (the slurry viscosity/the oil viscosity). The
model predicts that under certain conditions, the plugging of a well may not occur during restart even though hydrates would
be thermodynamically stable.

Conclusions
The hydrate kinetic model, CSMHyK-OLGA has been developed. This model provides the flow assurance engineer with a
valuable tool to predict the timescale of hydrate formation, and the tendency to form hydrate plugs. This kinetic model will
have applications in in developing new flow assurance strategies, including establishing operating procedures from transient
flow predictions such as restart and shut-in, and in flowline design.

The CSMHyK-OLGA kinetic model successfully predicts the hydrate formation rate in two flowloops (ExxonMobil, and
University of Tulsa flowloops) for three different oils using the same kinetic parameter (1/500 of intrinsic rate constant). The
reduced rate constant suggests that mass or heat transfer limitations limit hydrate formation in the flowloops, rather than
intrinsic kinetics.

Industrial application of the model has been implemented by Chevron to major capital projects. CSMHyK-OLGA provides a
worst case scenario of hydrate formation. Sensitivity analysis of the model demonstrated: (i) hydrate plugging can be
predicted on restart, (ii) the rate of heat removal from the pipeline often limits hydrate growth, (iii) reducing slip factors
causes hydrate accumulation.

Acknowledgements
The authors would like to first thank all DeepStar participants for funding this work in part. We would also like to recognize
the generous donation of flow loop time by Dr. Talley and Exxon Mobil, and Drs. Volk and DelleCase for the flow loop
experiments performed at the University of Tulsa. Thanks are also given to Dr. Xu from SPT Group for his help with the
OLGA developments, and Dr. Montesi of Chevron for his work in applying the model to flow lines and flow assurance
design.

References
1. Sloan, E.D. and Koh, C.A. Clathrate Hydrates of Natural Gases, 3
rd
Ed, CRC Press, Boca Raton, FL, 2008.
2. Sloan, E.D., Monograph of Hydrate Engineering, TX, 2000
3. Vysniauskas, A. and Bishnoi, P. R., (1983) Kinetics of Methane Hydrate Formation, Chem. Eng. Sci.1983; 38:
1061-1072
4. J. Boxall, S. Davies, K.T. Miller, C.A. Koh, E.D. Sloan, J. Creek, Hydrate Plug Formation Model, DeepStar Final
Report CTR 8201, 2007
5. Matthews P.N., Notz P.K., Widener M.W. and Prukop G. Flow Loop Experiments Determine Hydrate Plugging
Tendencies in the Field. Gas Hydrates: Challenges for the Future. Annals New York Academy of Sciences, 912,
330-338, 2000
6. Camargo R. and T. Palermo, (2002) Flow Properties of Hydrate Suspensions in Asphaltenic Crude Oil Proc. 4
th

Intnl. Hydrates Conf., Yokohama 2002; 880-885
7. J. Boxall, K. T. Miller, P. Rensing, C. Taylor, C. A. Koh, and E. D. Sloan, Hydrate Plug Formation Model on Flow-
Line Start-Up, DeepStar Final Report CTR 7201, 2005
0
500
1000
1500
2000
2500
3000
3500
4000
4500
5000
0 1000 2000 3000 4000 5000 6000 7000
Distance from Reservoir (ft)
P
r
e
s
s
u
r
e

(
p
s
i
a
)

O
R

V
i
s
c
o
s
i
t
y

R
a
t
i
o
0
2
4
6
8
10
12
14
16
V
e
l
o
c
i
t
y

(
f
t
/
s
)
Pressure
Viscosity Ratio
Gas Velocity
Liquid Velocity

Você também pode gostar