You are on page 1of 27

OTC 23965

Hydrate Prevention in Subsea Oil Production Dead-Legs


M. Haghighi, K. A. Hawboldt, and J. Cai, Memorial University of Newfoundland
M. A. Abdi and C. Sundgaard, Husky Energy

Copyright 2013, Offshore Technology Conference

This paper was prepared for presentation at the Offshore Technology Conference held in Houston, Texas, USA, 6–9May 2013.

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
A study was conducted to better understand the hydraulic and thermal behaviour of dead-legs within a stacked spool
arrangement designed to produce two well streams from a single manifold. The study included scenarios under different
rigid spool orientations and active line flow conditions. The ultimate goal of the study was to optimize inhibitor injection
procedures with respect to loss of inhibitor and contamination of the production fluids.
A Computational Fluid Dynamics (CFD) technique was employed to simulate the flow and thermal conditions within the
flow lines and dead-legs. The results of the simulations were validated with two sets of experiments and in two different
scales (sizes). The findings of the study suggest that the thermal and/or hydraulic status of the dead-legs during long shut-in
periods pose little risk of hydrate formation.

Introduction
A dead-leg is defined as an inactive portion of the pipe where the flow is stagnant or has very low velocity. This inactive pipe
is normally connected to an active pipe that carries the main flowing stream. Due to the low or stagnant flow created in dead-
legs, there is a risk of cool-down and hydrate formation. The stagnant conditions within the dead-legs could also promote
corrosion.
The formation of hydrates is a well known phenomenon in the oil and gas production and processing systems. Hydrates
can form under low temperature and high pressure when water is present in contact with gaseous or live crude oil and light
liquid hydrocarbons. The fluids present in the well-stream contain a certain amount of gas which is saturated with vaporized
water. Free produced or condensed water can collect in the low-points of the pipeline including pipe segments within dead-
legs. Hydrates can be formed as a solid block when the line is left inactive and can potentially plug the pipeline.
When hydrate formation happens, the hydrate plug in the pipeline may be either removed by depressurization or injection
of inhibitor. By depressurizing one or even two sides of the plug, the hydrate plug starts to melt and can potentially move
with high velocity, which could cause severe damage to subsea components. The formation of hydrates in the rigid spools
connecting the subsea trees to the main subsea production manifolds could result in complex consequences. The manifold-
subsea tree interconnecting spool is connected to the manifold from one side and can be relatively easily depressurized;
however the tree side of the spool cannot be depressurized readily and may require costly subsea intervention and an
extended loss of production. Addition of hydrate inhibitor will reduce the melting temperature and reduce the thermodynamic
potential for hydrate formation. Flushing dead-legs with inhibitors such as methanol can protect them from hydrate formation
during shut-in periods.
The conventional manifold-subsea tree interconnecting spools can be readily isolated from both sides using tree and
manifold valves and therefore there is no risk for methanol loss to the active line. The spools therefore remain fully flushed
with methanol and protected from hydrates during long shut-in periods. However, when installation restrictions favor a
2 OTC Error! Reference source not found.

stacked spool, where two trees are connected to the main manifold connection hub, the individual spools can only be isolated
from the well side and there is a risk of methanol loss during well testing of one of the two wells.
Traditionally the tree-manifold spool attached to the inactive well is flushed with methanol during long shut-in periods.
When only one well is shut-in and the other is producing, the methanol in the dead-leg can be lost to the active line and
therefore there is a potential for the dead-leg to be exposed to production fluids under low subsea temperatures and hydrate
formation conditions. The continuous injection of methanol can be costly and may lead to loss of methanol in the active line
which in turn can also give rise to further contamination of produced water released to the ocean. Therefore, it is important to
investigate the loss of methanol from the inactive spool and potential for hydrate formation in the interconnecting spools
during long-term shut-in periods. The frequency and quantity of methanol injection required and the revision to current
operating procedures should also be determined.

Literature Review
There have been several studies on multiphase flow in straight pipelines and pipelines with T-junctions. However, there is
little published work regarding a mass transfer model for determining hydrate inhibitor dilution in the production line dead-
legs. The impact of the thermal and hydraulic status of the active flowline on the dead-leg’s cool-down time under various
gas-to-oil ratios (GORs) in the active line is also not addressed in previous studies.
Habib et al. (2005) investigated the effect of the dead-leg geometry and orientation on the velocity field and oil-water
separation in the dead-leg. Horizontal and vertical dead-legs with different diameter-to-length ratios were investigated and
the mathematical formulation for the calculation of the fluid-flow field was also established. In their studies, the fluid-flow
model was based on the time-averaged governing equations of 3D turbulent flow. The algebraic slip-mixture model was
utilized for the calculation of the two immiscible fluids (water and crude oil). The results showed that the size of the stagnant
fluid region increases with the increase of length-to-diameter ratio (L/D). For the case of a vertical dead-leg, it is found that
the region of the dead-leg close to the header is characterized by circulating vortical motions for a length of L=3D while the
remaining part of the dead-leg is occupied by a stagnant fluid. In the case of a horizontal dead-leg, the region of circulating
flow extends to 3–5 times D. The results also indicated that the water volumetric concentration increases with the increase of
L/D and is influenced by the dead-leg orientation. A visualization experiment was carried out to validate the calculation
procedure. However, thermal behaviour and heat transfer in the dead-leg were not included in this study. Based on this work,
Andersen et al. (2007) studied heat transfer and flow pattern in two geometries under different velocities for hydrate control
purposes in a natural gas application. The effect of buoyancy in the vertical dead-leg was also investigated. The results show
that the vertical dead-leg has a more abrupt temperature profile outwards in the dead-leg than the horizontal one and the
temperature crosses the hydrate equilibrium temperature (HET), where hydrates would start to form. This simulation was
only limited to un-insulated dead-legs. In the case of a dead-leg with insulation, heat loss along the pipeline wall decreases
and therefore the temperature can potentially be maintained higher and beyond the range of hydrate equilibrium temperature
for the same distance from the main line. However, no research has been published on the effects of hydraulic and thermal
behaviour of the active line on methanol loss from the dead-leg. This study aims at investigating the hydraulic and thermal
behaviour of dead-legs under the flowing crude oil production conditions and insulated flowlines and the impact of dead-leg
orientation and active flowline conditions on methanol loss in the dead-leg.

Objectives
Given that the hydraulic and thermal characteristics of the dead-leg are a function of the thermal and hydraulic state of the
active flowline, the objectives of this project are:
 To review and make a summary of literature on heat transfer and flow patterns due to natural convection with
focus on horizontal and vertical dead-legs.
 To investigate the hydraulics and thermal behaviour of stacked spool dead-legs; both vertical and horizontal
orientations will be investigated.
 To study the potential for fluid displacement in the dead-leg.
 To set up and perform Computational Fluid Dynamics (CFD) study of flow in the dead-leg. The tool for the CFD
study is the commercial package Fluent from ANSYS 13.0.
 To validate the CFD predictions through bench and pilot scale experiment set-ups.
OTC 23965 3

Study basis
The CFD technique was used to simulate flow conditions during shut-in operations for a stacked spool configuration as
shown in Figure 1. The geometry used for the simulations is based on the example spool geometry shown in Figure 1. The
internal diameter (ID) of the pipes is assumed to be 0.1244 meters and the insulation is modeled by an overall heat transfer
coefficient (OHTC) of 3.7 W/m2.K based on the inside diameter (ID) of the pipe. The initial simulations were based on
simplified flow conditions as described in Table 1.
Table 1 – Initial simulations specifications
Gas phase Methane
Liquid Phase Water
Operating temperature 100 ˚C
Operating Pressure 60 bar g
3
Total flow rate 1,500 m /day
3 3
GOR 120 Sm / Sm
Dead-leg Initial Conditions Filled with methanol at ambient temperature (-1.8 ˚C)

Figure 1 – Geometry of the stacked rigid spool studied.

These conditions were modeled by an inlet mass flow of 17.3 kg/s water and 1.35 kg/s methane at the specified
temperature. The pressure at the location of the T-junction is set to 6,101,325 Pa and pressure profile throughout the spool is
determined by solving the conservation equations. The Soave-Redlich-Kwong (SRK) equation of state was used to predict
4 OTC Error! Reference source not found.

the non-ideality in the gas phase. Constant properties were used for the liquid phase. More realistic conditions were used for
the final simulations as described in Table 2.

Table 2 – Final simulation specifications


Gas phase Real gas mixture properties
Liquid Phases Real fluid mixture properties
Operating Temperature 100 ˚C
Operating Pressure 50 barg
3
Total Liquid Flow Rate 1,200 m /day with 90% water cut
3 3
GOR 1,000 Sm / Sm
Dead-leg Initial Conditions Filled with methanol at ambient temperature (-1.8 ˚C)

These conditions were also modeled by a mass flow inlet, with flow rates of water, oil, and gas at 12, 1.15, and 1.18 kg/s
respectively. Temperature and pressure were adjusted accordingly.

CFD Models and Governing Equations


There are different built-in models within CFD package Fluent 13.0 (ANSYS Fluent Theory Guide, 2010 and ANSYS Fluent
User’s Guide, 2010) to simulate multiphase flow. The simplest of these is the Volume of Fluid (VOF) model. VOF is the
model of choice for simple multiphase flow where mixing or separation is not of the main interest. The most advanced model
available is the Eulerian model which solves a different set of conservation equations for each phase and is capable of
modeling complex multiphase flows. This is the model of choice for more complicated and advanced simulations in this
study. A brief summary of the modeling techniques for each of these models is presented here.

Volume of Fluid (VOF)


In the volume of fluid (VOF) model, a single set of conservation equations is shared by the phases and the volume fraction of
each of the phases is tracked in each computational cell throughout the domain. These conservation equations can be solved
using the appropriate jump boundary conditions at the interface. However, the interface between the different phases does not
necessarily remain stationary. As a result, imposing boundary conditions at an interface becomes a very complicated moving
boundary problem. To avoid this problem, the VOF model directly determines the motion of all phases and deduces the
motion of the interfaces indirectly from the results. Direct tracking of the motion and the deformation of the phase interfaces
is thus avoided. All the interfacial forces, therefore, have to be replaced by smoothly varying volumetric forces. If one is in
particular interested in the geometry of the interface and the flow patterns near the interface, the VOF model is an interesting
model to use. Some interface-related forces, such as surface or adhesion forces, can also be modeled accurately using this
modeling approach. For each additional phase that is added to the model, a variable is introduced: the volume fraction of the
phase in the computational cell. In each control volume, the volume fractions of all phases sum up to unity. All control
volumes must be filled with either a single fluid phase or a combination of phases. The VOF model does not allow for void
regions where no fluid of any type is present. The values for all variables and properties are shared by the phases and
calculated as volume-averaged values, as long as the volume fraction of each of the phases is known at a given location.
Thus, the variables and properties in any given cell are either purely representative of one of the phases or representative of a
mixture of the phases, determined by the volume fraction contributions. In other words, if the kth fluid’s volume fraction in
the cell is denoted as  k , then the following three conditions are possible:

 k  0 : the cell does not contain the kth fluid;


 k  1 : the cell is full of the kth fluid;
0   k  1: the cell contains the interface between the kth fluid and one or more other fluids.
Based on the local value of  k , the appropriate properties and variables will be assigned to each control volume within the
domain.
In the VOF approach, the participating fluids share a single set of conservation equations. The governing equations can
therefore be written as (ANSYS Fluent Theory Guide, 2010):
OTC 23965 5

Conservation of mass:

(  )    ( U )   Sk (1)
t k

Conservation of momentum:

( U )    ( UU )       g  F (2)
t
The first term on the left-hand side of equation (2) is the rate of increase in momentum per unit volume; the second term
represents the change in momentum per unit volume, caused by convection. The first term on the right-hand side represents
molecular contributions, which include pressure and viscous force per unit volume. The last two terms on the right-hand side
represent the gravitational force per unit volume and any other external force. It is straightforward to apply such conservation
equations to single-phase flows. In the case of multiphase flows, the volume fraction of each fluid  k  is calculated by
tracking the interface between different phases throughout the solution domain. Tracking of the interfaces between N
different phases present in the system is accomplished by solving continuity equations of the phase volume fractions for N−1
phases. For the kth phase, the conservation of mass has the following form:
k
   uk k    J k  k (3)
t

 k   k  k , J k  0,  Mass transfer rate


 ( k  k )  n
 uk  k  k )  S k   (m pq  m qp ) (4)
t p 1

The heat transfer can be calculated by the energy equation for the mixture as follows:
 
(  E )    (v (  E  p ))  S h    ( K eff T )  S h (5)
t
Where
n


q 1
q  q Eq
Ek  n
(6)

q 1
q q

Where E q for each phase is based on the specific heat of that phase and the shared temperature. The properties  and Keff
(effective thermal conductivity) are shared by the phases. The source term Sh contains contributions from radiation, as well as
any other volumetric heat sources.

Eulerian Model
The Eulerian multiphase model in ANSYS Fluent allows for the modeling of multiple separate, yet interacting phases. The
phases can be liquids, gases, or solids in nearly any combination. With the Eulerian multiphase model, the number of
secondary phases is limited only by the memory requirements and convergence behavior. Any number of secondary phases
can be modeled, provided that sufficient memory is available. The description of multiphase flow as interpenetrating continua
incorporates the concept of phasic volume fractions, denoted here by αq. Volume fractions represent the space occupied by
each phase, and the laws of conservation of mass and momentum are satisfied by each phase individually. The derivation of
the conservation equations can be done by ensemble averaging the local instantaneous balance for each of the phases
(Anderson and Jackson, 1967) or by using the mixture theory approach (Bowen, 1976).
The general conservation equations from which the equations solved by ANSYS Fluent are derived are presented in this
section, followed by the solved equations themselves.

Conservation of Mass
The continuity equation for phase q is
6 OTC Error! Reference source not found.

  n
( q  q )    ( q  q vq )   (m pq  m pq )  Sq (7)
t p 1

Where, vq is the velocity of phase q and m  pq characterizes the mass transfer from the pth to qth phase, and m pq characterizes
the mass transfer from phase q to phase p, and the user is able to specify these mechanisms separately. By default, the source
term Sq on the right-hand side of Equation 7 is zero.

Conservation of Momentum
The momentum balance for phase q yields
  
( q  q vq )    ( q  q vq ) 
t
(8)
 n      
  q p    q   q  q g   R pq  m pq v pq  m qp vqp )  ( Fq  Flift ,q  Fvm, q )
p 1

where  q is the qth phase stress-strain tensor:

  2 
 q   q q (vq  vqT )   q (q  q )  vq  (9)
3
  
Here μq and αq are the shear and bulk viscosity of phase q, Fq is an external body force, F lift , q is a lift force, F vm , q is a

virtual mass force, R pq is an interaction force between phases, and p is the pressure shared by all phases.

Experiments and Model Validation


In order to validate the simulation results, two sets of experiments were conducted. A series of bench scale experiments were
performed for a qualitative validation of the results of the application of the VOF and the Eulerian multiphase models in the
prediction of fluids displacement in and out of a dead-leg connected to an active line with two-phase flow. Additional
experiments were conducted on a larger scale set up for more robust and quantitative validation of the CFD models in the
prediction of fluids displacement and heat transfer. The following sections describe these experiments and their findings.

Bench Scale Experiment


As mentioned above, the VOF and the Eulerian models were selected to simulate the flow pattern and mass transfer in dead-
legs and built-in CFD codes were used to solve the governing equations. To validate the models, a bench scale model was
first set up and experiments were conducted. In the bench scale experiments (see Figures 2 and 3 for system arrangement),
transparent plastic tubes were used to visually observe the movement of fluids and qualitatively observe the fluid
displacement in the dead-leg. Six millimeter (inner diameter) tubing was used to create a scaled representation of the
geometry under investigation (Figure1). Methanol was stored in a storage tank and tinted with color, which made the
interface between fluids from the main flow line and methanol visually clear.
OTC 23965 7

Figure 2 – Horizontal dead-leg bench scale experiment.

Figure 3 – Vertical dead-leg bench scale experiment.

An air to water volume flow rate ratio of seven was tested in both vertical and horizontal dead-legs under room
temperature. The dead-leg was filled with methanol before each experiment and gas flow rates were measured using a
variable area rotameter. The specifications of the testing system are listed in Table 3.

Table 3 – Specifications of the fluids and equipment for bench scale testing system
Feed liquid water
Feed gas air
Air/water ratio 7-8 (volumetric)
Tube diameter 6 mm
Tube material Silicone
Scale ratio 1:25

The flow pattern in the main flow line falls in the dispersed regime for both vertical and horizontal dead-legs. It was
observed that in the vertical dead-leg methanol was quickly displaced by gas due to the impact of gravity and buoyancy.
Very little water broke into the vertical part connected to the active line, up to a length of approximately 5D.
8 OTC Error! Reference source not found.

Unlike the vertical dead-leg, the displacement of methanol in the horizontal dead-leg was much slower. Gas displacement
was observed in the beginning where the driving force mainly depended on gas displacement. As gas gradually displaced
methanol, water intruded into the horizontal line connected to the active flowing vertical line and the interface between water
and methanol mainly stayed within the horizontal line for a period of one hour. The driving force for mass transfer of
methanol changed from mainly gas displacement to diffusion and gas displacement, but methanol loss in the dead-leg was
much less and occurred more slowly compared to the vertical dead-leg.
CFD simulations were subsequently developed for both horizontal and vertical dead-leg tests with similar properties and
geometries to those of the actual bench scale experiments. The simulations for the horizontal dead-leg showed similar results
in the depletion of methanol from the leg. Methanol was displaced from the horizontal section as seen in Figure 4. The
location of the methanol/water interface remained in the horizontal section, unlike the tests conducted with larger scale test
system where the methanol was completely displaced with the air in the horizontal section (explained below). This was
attributed to the effect of surface tension in the small tubing and confirmed by two sets of simulations; one that did not model
this effect at all, and one that used a larger scale geometry and rendered it effectively negligible. In both of these simulations
the methanol depletion was advanced into the vertical section (Figure 5). The results of the simulations for the vertical dead-
leg are also in good agreement with those of the experiments (Figure 6).

Figure 4 – Methanol volume fraction contours of in the horizontal dead-leg simulation of the bench scale model
OTC 23965 9

Figure 5 – Methanol volume fraction contours showing depletion of methanol progressed into the vertical section as effects of
surface tension are ignored.

Figure 6 – Methanol volume fraction contours in the vertical dead-leg simulation of the bench scale model
10 OTC Error! Reference source not found.

Large Scale Experiments


The second stage of model validation was based on experiments with a flow loop including vertical and horizontal dead-legs
as shown in Figure 7. The flow loop is made with 2-inch steel pipes and is placed in a frame approximately 1 m by 2.3 m.
The dead-leg extends 0.7 m away from the main loop and it can rotate to take horizontal or vertical orientations (see Figures
8 and 9). Thermocouple probes are located at 12 positions along the leg with approximate distances of 51, 89, 127, 165, 203,
241, 292, 343, 394, 470, 546, and 622 mm from the centre of the active line T-junction. The thermal behaviour of the dead-
leg was examined by measuring the temperature profile along the dead-leg. Transparent Polyvinyl Chloride (PVC) pipes
were used for observation of the methanol loss from the dead-leg. A 1.5 kW centrifugal pump was used to circulate water at
up to 0.0036 m3/s (220 l/min) and a 3kW immersion heater was employed to heat the water in a 60 litre (15 gallon) tank to
replicate the hot reservoir. Table 4 summarizes the specifications of the pilot test set-up.

Figure 7 – Schematic of the pilot test flow loop.

Table 4 – Specifications of the fluids and equipment for large scale testing system.
Item Description
Liquid water
Feed gas air
Pipes 2” (carbon steel)
Transparent Pipe 2” (PVC)
Water tank 15 gal
Heater 3 kW
Thermometer 12 channel Bench top scanner
Temperature probes Type T thermocouples
Pump 1.5 kW centrifugal pump, cast iron housing and brass impeller
Flow meter 1.5” turbine flow meter
OTC 23965 11

Temperature
Rotatable
probes
dead leg

Water tank containing


immersion heater

Air from air


compressor

Pump

Figure 8 – The flow loop rig used for validation of experiments with the dead-leg at its vertical position.

Temperature
probes

Water tank

Figure 9 – Vertical (left) and horizontal (right) orientations of the dead-leg

Liquid Displacement
The displacement of liquids from the dead-leg was examined by visual observation of the displacement process through the
transparent PVC pipes (Figure 10). The experiments showed that the liquid in the vertical dead-leg is completely replaced by
the gas from the active flow line in less than a minute. For the horizontal leg, the liquids are replaced by the gas only in the
top region (vertical section immediately attached to the horizontal segment) of the pipe. The liquid in the lower parts of the
leg is replaced by liquids from the active line. This is completed in two stages. At first, the liquid in the beginning of the leg
is very quickly displaced due to the dynamics and the resulting convection of the fluids in the active line. The distance this
12 OTC Error! Reference source not found.

forced convection behaviour extends to is a function of the velocity and volume fraction of the liquid in the active line. The
second stage of the displacement takes place further in the leg where the velocity of the liquid is negligible. The mass transfer
mechanism at this stage is mainly diffusion and hence the displacement is much slower and takes a much longer time to
complete. The results of the CFD simulations matched the observations in the experiments well, as can be seen in Figures 11
- 14.

Transparent PVC
pipes

Methanol tank

Figure 10 – Flow loop with the transparent PVC dead-legs mounted on.
OTC 23965 13

Figure 11 – Volume fraction of water from the simulation based on the flow loop test conditions for the horizontal dead-leg.

Figure 12 – Snap shot of the experiment with two phase (air/water) flow, showing the location of air and water interface in the
horizontal dead-leg.
14 OTC Error! Reference source not found.

Figure 13 – Volume fraction of water from the simulation based on the flow loop test conditions for the vertical dead-leg.

Figure 14 - Snap shot of the experiment with two phase (air/water) flow, showing the location of air and water interface (dashed
lines) in the vertical dead-leg.
OTC 23965 15

Temperature Distribution
Experiments were performed with the dead-leg in both vertical and horizontal orientations. The temperature readings from
each of the 12 probes were registered for periods of up to 12 hours. In the case of the vertical dead-leg, the temperature rose
in the leg (initially filled with cold water) in about a minute as the water in the leg was gradually replaced by the air in the
active line. The steady-state condition was achieved in a fairly short period of time in which a relatively constant gradient
was observed along the entire length of the dead-leg. The temperature at various probe locations varied between the
temperatures of the flowing fluids at the T-junction to that of the ambient at the far end of the leg. It should be noted that the
far end of the leg was equipped with a brass valve which was not insulated and hence the temperature at that point was close
to the ambient temperature. The CFD model predictions also show the same characteristics (see Figure 15). In the case of the
horizontal leg, as it was observed in the displacement tests, the lower section of the leg remained filled with liquid, although
the initially cold liquid was replaced and/or heated by the hot flowing liquid. The result is again a rather constant gradient
throughout the entire length of the leg, but the lower extreme was nowhere near the ambient temperature. In fact, the
difference between the temperature of the leg at the T and its temperature at the far end was only about 2˚C (Figure 16). This
is mainly attributed to the continuous liquid (water) phase and liquid circulation throughout the horizontal leg as opposed to
air in the vertical leg.

70

60

50
Temperature (˚C)

40

Experiment
30
CFD

20

10

0
0 100 200 300 400 500 600 700
Distance from T (mm)

Figure 15 – Comparison of the experiment temperature readings and the results of the CFD model for the vertical dead-leg.
16 OTC Error! Reference source not found.

82

81.5

81
Temperature (˚C)

80.5

Experiment
80
CFD

79.5

79

78.5
0 100 200 300 400 500 600 700
Distance from T (mm)

Figure 16 – Comparison of the experiment temperature readings and the results of the CFD model for the horizontal dead-leg.

Simulation Results
The results of the CFD simulations are presented in two categories. The first is the results of the preliminary simulations with
water and methane as flowing fluids. These simulations were based on the exact same modeling techniques whose validity
was examined against experimental data. The second group of results is predictions from the CFD model that was tailored to
the example subsea production flow conditions. In this model the production conditions are more difficult to simulate exactly
and therefore are better suited to identify trends rather than exact values. However, the good agreement of these results to
those of the first group is an important indicator of their soundness and reliability.

Preliminary models (water – methane)


These simulations used the following simplifying assumptions in the flow conditions:
 Only water was assumed present as the single liquid phase. This means that the volume fraction of oil in the actual
flow conditions is ignored.
 The gas phase is modeled as pure methane while the actual gas phase is a combination of mainly methane and
several other hydrocarbon gases, as well as nitrogen and carbon dioxide.
 Gas-to-water ratio is about 120 Sm3/Sm3 whereas in the actual flow conditions it may be as high as 1,000 or more.
Note that the assumption of 100% water is a conservative in the context of this study. In reality wells could produce until
the volume fraction of water reaches about or above 95%. The higher the volume fraction of water is in the active line, the
higher will be the potential for water to intrude into the dead-leg. Also it is assumed that methane will be representing the
associated gas in the rigid spools. The molecular weight of the associated gas in the spool will be a function of spool
temperature and pressure and could vary between 17 and 19 as compared to methane molecular weight of 16. Given the
physical properties of gas including heat capacities, thermal conductivity, density and viscosity are marginally impacted by
the gas molecular weight (±10-15%), the assumption of methane only in the spool is not going to be too far from reality.
Also in this method, the thermo-physical properties of methane and water were taken from the Fluent’s materials library,
with the exception of methane density which was calculated based on the SRK equation of state.
OTC 23965 17

Vertical Dead-leg
The simulations of this situation show that the methanol initially in the dead-leg is quickly displaced with methane from the
active line and nearly no water intrudes in the dead-leg. This is also in agreement with the experiments conducted with the
flow loop. Figures 17 and 18 show the methanol and water volume fractions respectively at steady state conditions, 20 hours
into the simulation.

Active spool

Dead-leg

Figure 17 – Steady state methanol volume fraction contours showing the displacement of methanol from the vertical dead-leg.
18 OTC Error! Reference source not found.

Active spool

Dead-leg

Figure 18 – Steady state water volume fraction contours, showing that water has not intruded into the vertical dead-leg.
OTC 23965 19

Horizontal Dead-leg
The results of the simulations show that methanol is again displaced out of the dead-leg. Most of the pipe in the leg is filled
with gas, but the first horizontal section (the first 1.6m) is occupied by water. The existence of water presents a potential for
hydrate formation, but the temperature in that region remains above the example hydrate equilibrium curve shown in Figure
19 and hence poses little risk for hydrate formation. Figures 20 and 21 show the volume fraction contours of methanol and
water and Figures 22 to 24 show the temperature profiles in the dead-leg at steady state conditions (20 hours into the
simulation).

Figure 19 – Hydrate equilibrium curve for methane.


20 OTC Error! Reference source not found.

Active spool

Dead-leg

Figure 20 – Steady state methanol volume fraction contours showing depletion of methanol from the horizontal dead-leg.

Active spool

Dead-leg

Figure 21 – Steady state water volume fraction contours showing the intrusion of water into the horizontal section of the dead-leg.
OTC 23965 21

Active spool

Dead-leg

Figure 22 – Temperature (K) contours showing that the initially cold dead-leg is warmed up by the intrusion of hot fluids from the
active line.

100
98
96
94
Temperature (C)

92
90
88
86
84
82
80
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
Length (m)

Figure 23 – Plot of steady state temperature profile at the water/methane interface in the horizontal dead-leg.
22 OTC Error! Reference source not found.

90

85

80
Temperature (C)

75

70

65

60

55

50
0 0.5 1 1.5 2 2.5
Length (m)

Figure 24 – Plot of steady state temperature profile at the centre of the pipe in the vertical section of the dead-leg.

Complete models (with actual fluids flowing)


In this group of simulations actual fluids were flowed through the flowlines as opposed to assumed water and methane in the
previous simulation runs. The physical properties of the fluids including oil, gas, and water were input into the simulations as
polynomial functions (fitted to look-up tables prepared by PVTSIM software) where the required physical property
information could be available to the simulations. As a result, the simulations were more complicated and required more
computer processing power and/or longer computation time to complete. Moreover, the introduction of fluid properties that
are not of a pure material requires unconventional modeling strategies that may result in uncertainties. The initial approach
was using polynomial functions of temperature to define different thermo-physical fluid properties. While this technique
works well for all other properties (such as viscosity and specific heat capacity), it presents certain problems with density due
to the absence of correlation with respect to pressure. A second approach was devised to overcome this complication; the
SRK equation of state was used in conjunction with pseudo-critical properties (derived via Kay’s mixing rule) that would fit
the density profile of the reservoir vapours within the flow conditions. All other properties were defined as polynomials fitted
to the data provided by generated by PVTSIM. The results of these simulations are very much similar to those of the
simplified models, except for the existence of a third phase (oil phase) in the active flow line. This specifically stands out in
the case of the horizontal dead-leg. As seen in Figures 25 to 30, liquids intrude into the dead-leg in the same fashion observed
in the simplified model. Water, as expected, occupies the lower section of the horizontal dead-leg while oil and gas fill the
middle and top sections, respectively (see Figures 25 - 27). The temperature in the dead-leg also rises similar to the simplified
models and due to the intrusion of hot fluids from the active line. The temperature of the spool in the liquid filled sections
remains well above hydrate formation temperature shown in Figures 28 to 30. Hydrate formation temperature for the
maximum pressure (5 MPa) ranges between 13 to 16˚C depending on the fluids GOR.
OTC 23965 23

Active spool

Dead-leg

Figure 25 – Steady state gas phase volume fraction contours from the simulation of the horizontal dead-leg with real fluids
properties.

Active Spool

Dead-leg

Figure 26 – Steady state oil phase volume fraction contours from the simulation of horizontal dead-leg with real fluids properties.
24 OTC Error! Reference source not found.

Active spool

Dead-leg

Figure 27 – Steady state water volume fraction contours from the simulation of the horizontal dead-leg with real fluids properties.

Active spool

Dead-leg

Figure 28 – Steady state absolute temperature (K) contours from the simulation of the horizontal dead-leg with real fluids properties,
showing the warming up of the dead-leg due to intrusion of hot fluids from the active line.
OTC 23965 25

100
99.5
99
98.5
Temperature (˚C)

98
97.5
97
96.5
96
95.5
95
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
Length (m)

Figure 29 – Plot of steady state temperature profile at the centre of the pipe in the horizontal dead-leg with actual production fluids
flowing in the active spool.

70

65

60
Temperature (C)

55

50

45

40

35

30
0 0.5 1 1.5 2 2.5
Length (m)

Figure 30 – Plot of steady state temperature profile at the centre of the pipe in the vertical section of the dead-leg with actual
production fluids flowing in the active spool.

Conclusion
Based on the findings of this study including the results of the CFD simulations and observations of the experiments
(summarized in Tables 5 and 6), it is concluded that the risk of hydrate formation in a dead-leg formed in any of the spools
due to a shut-in operation while the other spool is producing, is minimal. In a vertical dead-leg the CFD simulations indicated
that under actual operating conditions water does not intrude in the dead-leg and hence the possibility of hydrate formation is
unlikely. In a horizontal dead-leg the CFD simulations indicated that under the actual operating conditions, water does settle
in the dead-leg. However, the temperature in the dead-leg rises with the intrusion of water and other fluids from the active
line into the stagnant spool and remains at a level much beyond the hydrate equilibrium envelope. As a result, the risk of
hydrate formation in this configuration is also considered minimal.
26 OTC Error! Reference source not found.

Table 5 – Summary of the results of the experiments and CFD model validation
Experiment Measurements CFD Results

Flowing Temperature 61.5 61.5

Temperature in at the beginning of the leg 61 61


Horizontal
Dead-leg (˚C) in the middle of the leg 42 45

at the far end of the leg 28 27

Flowing Temperature 81.5 81.5

Temperature in at the beginning of the leg 81.5 81.5


Vertical Dead-
leg (˚C) in the middle of the leg 80 80.4

at the far end of the leg 79 79.5

Table 6 – Summary of the CFD simulations using the actual spool geometry

Gas Phase Methane Methane Real mixture Real mixture


Flowing
Fluids
Liquid Phase(s) Water Water Real mixture Real mixture

Operating Temperature (˚C) 100 100 100 100

Operating Pressure (bar g) 60 60 50 50

Dead-leg Orientation Vertical Horizontal Vertical Horizontal

Fluids Gas Phase Yes Yes Yes Yes


Intrusion in
Dead-leg Liquid Phase No Yes No Yes

Min Temperature at the water-


reservoir fluid (gas or oil) interface in 100 82 100 95.5
dead-leg (˚C)

Hydrate Equilibrium Temperature (˚C) 6-8 6-8 13-16 13-16

Temperature at the gas-methanol


43 40 47 47
interface (˚C)

Steady state time (hours) 8 10 8 11

Potential for hydrate formation unlikely unlikely unlikely unlikely

Acknowledgements
The financial support provided by Husky Energy and MITACS Accelerate is greatly appreciated.
OTC 23965 27

Nomenclature
D = Diameter
E = Enthalpy
GOR = Gas to oil ratio
GWR = Gas to water ratio
ID = Internal Diameter
L = Length
OD = Outer Diameter
OHTC = Overall Heat Transfer Coefficient
p = Pressure
S = Source term
SRK = Soave-Redlich-Kwong equation of state
T = Temperature
U = Velocity (shared between phases)
v = Velocity (of each individual phase)
VOF = Volume of Fluid model
α = Phase volume fraction
μ = Viscosity
ρ = Density

 = Strain Tensor

References
Anderson, H. (2007). Computational study of heat transfer in subsea deadlegs for evaluation of possible hydrate formation
(Master’s Thesis), Telemark University College, Norway.
Anderson, T.B. & Jackson, R. (1967) A Fluid Mechanical Description of Fluidized Beds, Industrial & Engineering Chemistry
Fundamentals, 6(4), 527–539,
ANSYS FLUENT Theory guide – Release 13.0, (2010).
ANSYS FLUENT User’s Guide – Release 13.0, (2010).
Bowen, R.M. (1976) Theory of Mixtures. New York: Academic Press.
Habib, M. A.; Badr, H. M.; Said, S. A. M.; Mokheimer, E. M. A.; Hussaini, I.; &Al-Sanaa, M. (2005). Characteristics of flow
field and water concentration in a horizontal deadleg, Heat and Mass Transfer, 41(4), 315-326

You might also like