Deepstar - Multiphase Flow
Deepstar - Multiphase Flow
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
TABLE OF CONTENTS
POLICY STATEMENT
policy.pdf
1.0
INTRODUCTION
Sec1.pdf
2.0
Sec2.pdf
3.0
DESIGN PROCESS
Sec3.pdf
4.0
Sec4.pdf
5.0
MULTIPHASE FLOW
Sec5.pdf
6.0
Sec6.pdf
7.0
THERMAL MODELING
Sec7.pdf
8.0
TRANSIENT OPERATIONS
Sec8.pdf
9.0
HYDRATES
Sec9.pdf
10.0
PARAFFIN WAXES
Sec10.pdf
11.0
ASPHALTENES
Sec11.pdf
12.0
EMULSIONS
Sec12.pdf
13.0
SCALE
Sec13.pdf
14.0
EROSION
Sec14.pdf
15.0
CORROSION
Sec15.pdf
16.0
SOLIDS TRANSPORT
Sec16.pdf
17.0
SLUGGING
Sec17.pdf
18.0
SLUGCATCHER DESIGN
Sec18.pdf
19.0
PIGGING
Sec19.pdf
20.0
OTHER OPERATIONS
Sec20.pdf
21.0
H-0806.35
01-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
22.0
Sec22.pdf
23.0
Sec23.pdf
24.0
Sec24.pdf
25.0
Sec25.pdf
APPENDICES
SECTION 9 APPENDIX A: GAS HYDRATE STRUCTURES, PROPERTIES, AND
HOW THEY FORM
Sec9 App A.pdf
SECTION 9 APPENDIX B:
PROGRAMS
H-0806.35
ii
01-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
POLICY STATEMENT
H-0806.35
1-Dec-00
1.0
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
INTRODUCTION
The Flow Assurance Design Guide (FADG) sets forth basic engineering requirements
and recommended practice deemed necessary for the reliable and cost effective design
and operation of multiphase production systems. Because flow assurance is a multidiscipline activity, the FADG addresses each discipline and explains how each fit in the
overall design process. The major flow assurance technologies covered in the guide are:
H-0806.35
1-1
1-Dec-00
2.0
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
2-1
1-Dec-00
3.0
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
DESIGN PROCESS
This section describes the design methodology or process that the flow assurance
engineer follows in developing a successful, cost effective subsea production system and
its operating philosophy. The flow assurance design methodology flow chart is presented
in Figure 3-1 and forms the basis for the discussion on the design process. Links to the
relevant sections in the Flow Assurance Design Guide are also provided.
As illustrated in the design methodology chart, the flow assurance design process
involves several major steps:
Each step can be addressed individually; however, all steps will be considered
collectively because they are inter-related. The chart shows some of the considerations
and/or decisions involved with each step. For purposes of illustration, design process
steps are generally shown to be sequential. However in practice, several of the steps will
need to be addressed simultaneously.
The flow assurance design process starts early in the field development effort, potentially
even before any wells have been drilled when the types and amounts of reservoir fluid
samples are specified. The general sequence begins with the development of the design
basis and then the thermal-hydraulic design and assessment of fluid behavior. During the
thermal- hydraulic design phase, the flow assurance engineer will begin to interface with
other design engineers, such as pipeline/flowline and facilities engineers. In what is
typically a parallel effort, the flow assurance engineer will interface with the subsea
mechanical designers and other engineers, will develop operating strategies, and will
assist in determining host facility requirements. An over-riding consideration in the
design process is system economics and risk management.
H-0806.35
3-1
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The design process will be iterative due to inevitable changes in the design basis, interim
results during the design, changes in system economics, and other changes. Such
iterations are indicated in the design methodology flow chart at the decision points in
which no would be the answer.
The flow assurance design process involves multiple technical interfaces.
engineering, completions engineering, pipeline/flo wline mechanical design,
controls engineering, facilities engineering, and operational personnel will
with flow assurance during the design process. The numerous interfaces
effective project management.
3.1
Reservoir
subsea and
all interact
necessitate
Design Basis
The first major effort in the design process is to establish the design basis. The flow
assurance engineer will be directly involved in terms of determining and documenting the
fluid characteristics, in terms of both PVT behavior and the potential for solids formation.
For the other aspects of the design basis, such as reservoir behavior, site characteristics,
and host facilities, the flow assurance engineer will need to ensure that the data needed
for the flow assurance analyses are included in the design basis. Thus the flow assurance
engineer will need to interface with those responsible for reservoir engineering,
metoceanic data, bathymetry, and surface facilities. These interfaces will continue
throughout the project. It is important to note that the design basis will need to be built
with conservatism to offset poor or missing data.
This step in the design process assumes that fluid samples have been collected. A
substantial amount of laboratory work may be required to determine the characteristics of
the fluid samples. Standard PVT measurements should be performed on the fluids, and
then fluid characterizations should be developed for use in thermal-hydraulic and other
modeling (reservoir and process). Section 4 discusses PVT behavior and fluid
characterization. The fluids should also be tested for potential solids formation such as
wax and asphaltenes.
3.2
Thermal-Hydraulic Design
The thermal- hydraulic design effort evaluates the lifecycle performance of the entire
production system. All parts of the system and all interfaces must be considered
throughout the operating lifetime of the development. This effort also should include
assessment of the potential for flow reductions due to solids formation.
H-0806.35
3-2
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
At the beginning of this step, basic design and operating principles should be set.
Examples include methods to be used to keep the production system out of the hydrate
formation region. For oil systems this could mean insulation. For gas system this would
require chemical inhibition. Another example would be to establish a lower limit on well
production rate and/or to use insulated tubing to prevent wax deposition in wellbores
during normal operation. This could be extended to the flowline, or wax may be
managed in the flowline with pigging and chemical injection. Such principles help guide
the flow assurance engineer through the design process; however, these principles should
be continuously evaluated in light of system operability and economics.
3.2.1
Hydraulic Modeling
Most system design attributes can be set on the basis of steady state analyses. Steadystate hydraulic models are used to determine the diameters for production tubing,
production flowlines, injection flowlines, and export pipelines. Criteria for line sizing
include pressure constraints, flow rates, and erosional velocity limits. As part of the line
sizing exercise and hydraulic assessment, changes in parameters such as production rates,
water cut, and GOR during the field life need to be evaluated. Artificial lift may also be
considered. Operating pressures will be calculated. Sections 5, 6, and 14 deal with
multiphase flow, line sizing, and erosion.
3.2.2
Thermal Modeling
Thermal modeling is typically combined with hydraulic modeling, thus thermal- hydraulic
modeling. Operating temperatures are calculated as a function of insulation level and
other parameters initially via steady state analysis. Section 7 covers thermal modeling.
3.2.3
H-0806.35
3-3
1-Dec-00
3.2.4
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
3.3
3.4
Operating Strategies
Operating strategies must be consistent with the system design and should be adaptable to
suit new circumstances in the event that fluid characteristics or other system
characteristics are found to be significantly different from those in the design basis.
Development of operating strategies is presented in Section 21.
3.5
H-0806.35
3-4
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
20, and 22 cover slugcatcher design, pigging, other operations, and host facility
requirements.
3.6
System Economics
There are numerous detailed design and manufacture activities and considerations that
are implicitly lumped into the Assess System Economics step. Section 23 covers system
economics and risk management.
H-0806.35
3-5
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 3-1: FLOW ASSURANCE DESIGN PROCESS
ESTABLISH DESIGN BASIS
Reservoir Behavior
as f(t)
Productivity Index
Production Profiles
Pres. vs. Depletion
Temperature
Fluid Behavior
PVT Characterization
Hydrates
Wax
Asphaltenes
Flowline
Routing
Bathymetry
Seabed Temp.
INTERFACE WITH
MECHANICAL DESIGN
Host Facilities
Separator Pres.
Acceptable Arrival
Temp.
OPERATING
STRATEGIES
Model
Wells
Model
Flowlines
Model
Wells
Select Tubing
and Flowline
Diameters and
# of Flowlines
No
Plateau and
EOL Conditions
Satisfied?
Hydrates
Wax
Asphaltenes
Scale
Model
Flowlines
Prediction
Control
Remediation
Select Tubing
and
Flowline
Insulation
ThermalHydraulics OK?
No
HOST FACILITY
REQUIREMENTS
ASSESS
TRANSIENT
THERMALHYDRAULICS
Yes
No
ThermalHydraulics and
Fluid Behavior
OK?
Yes
Reservoir, Flow
System, and Host Design
Compatible?
Yes
Processing Capabilities
Processing Pres./Temps.
Metering
Storage Volumes
Export Requirements
PCS
Chemical Injection Pumps
Chemical Storage
Flare Requirements
Utility & Emergency Power
Surge/Slug Volumes
Surge/Slug Control
Startup / Warmup
Shutdown / Cooldown
Turndown / Ramp-up
Depressurization
Slugging
Yes
No
Procedures
Valve Sequences
Pump Sequences
Chemical Injection
Rates
Activity Durations
ASSESS FLUID
PHASE
BEHAVIOR
THERMAL DESIGN
ASSESS SYSTEM
ECONOMICS
CAPITAL COST
OPERATING COSTS
NET PRESENT VALUE
No
System
and Economics
Optimum?
Yes
DONE
H-0806.35
3-6
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
4.0
4.1
Introduction
Modeling of oil and gas production, processing, and transportation system requires
knowledge of how the fluid behaves with changes in temperature and pressure. This
modeling work will require not only fluid properties (densities, viscosities, heat
capacities) but also the phase behavior of the fluids.
Multi-component phase behavior is a complex phenomenon, which requires accurate
determination if two-phase pressure loss, hold-up and flow regime are to be determined
with any degree of confidence.
The phase behavior will determine the vapor-liquid split and the thermodynamic
properties of each of the phases present, and it is important for the designer of such a
system to have a knowledge of what form this equilibrium takes, and how it may change
in different parts of the pipeline. Since it is expected that both temperature and pressure
will fall as fluids flow along the pipeline, it is possible that either condensation or
evaporation will take place within the pipe. This can have a significant effect on liquid
holdup and hence pressure drop. It also means that the CGR (or GOR) can vary
significantly depending on whether it is based on pipeline inlet or outlet conditions, and it
is therefore important to make it clear under what conditions it has been calculated.
In practice, experimentally determined phase behavior is often limited and one has to
employ some method of prediction. There are two approaches commonly employed in
the prediction of hydrocarbon phase behavior. These are the black oil method, which
assumes that only two components, i.e. gas and liquid, make up the mixture, and the socalled compositional approach, in which each hydrocarbon component is taken into
account. The methods have their own relative merits and are discussed in this section.
This section also addresses fluid sampling. Without appropriate sampling techniques,
sample handling, and analysis methods, the predictive methods used in modeling of the
production and processing of reservoir fluids will be in error.
H-0806.35
4-1
1-Dec-00
4.2
Reservoir Fluids
4.2.1
Phase Behavior
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Reservoir fluids are often described in terms of their phase behavior, which can be
defined as the relationship between the fluid phases (usually the gas and the
oil/condensate) and how the phases change with variations in temperature and pressure.
Single Component Phase Behavior
In describing phase behavior, a system consisting of a single, pure substance is
considered first. Such a system behaves differently from systems made up of more than
one component. A phase diagram (or phase envelope) is a plot of pressure versus
temperature showing the conditions under which the various phases will be present.
Figure 4.2-1 shows a phase diagram for a single, pure substance.
4000
CRITICAL
POINT
Pressure
3000
LIQUID
SOLID
2000
VAPOR PRESSURE
CURVE
1000
TRIPLE
POINT
VAPOR
0
-50
50
100
150
200
Temperature
H-0806.35
4-2
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
4-3
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
7000
6000
CRITICAL
POINT
BUBBLE
POINT
CURVE
5000
Pressure (psia)
CRICONDENBAR
4000
LIQUID
99
MOLE %
QUALITY
LINES
3000
95
MOLE %
CRICONDENTHERM
90
MOLE %
80
MOLE %
2000
1000
VAPOR
0
-100
100
200
300
400
500
600
700
800
900
Temperature (F)
H-0806.35
4-4
1-Dec-00
1000
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
bubble point. The pressure at which the first gas is formed is the bubble point pressure.
As the pressure is decreased below the bubble point pressure, more and more gas appears.
Critical Point
As can be seen when comparing Figures 4.2-1 and 4.2-2, the definition for the critical
point for a single component is not the same as that for a multiple component mixture. A
rigorous definition of the critical point is that it is the point at which all properties of the
liquid and the gas become identical.
Cricondentherm and Cricondenbar
The highest temperature on the two-phase envelope is called the cricondentherm. The
highest pressure on the two-phase envelope is called the cricondenbar. These are
illustrated on Figure 4.2-2.
Quality Lines
Another feature in the two-phase envelope are quality lines. These lines indicate curves
on constant vapor or liquid quantities within the two-phase region. In Figure 4.2-2 there
are quality lines for 99, 95, 90 and 80 mole percent vapor. The quality lines all converge
at the critical point.
Retrograde Condensation
For many multiple component mixtures a phenomenon called retrograde condensation
can occur. If the mixture is at a pressure greater tha n the cricondenbar and at a
temperature greater than the critical temperature, it will be single-phase gas. If the
pressure is decreased isothermally, the dew point curve will be crossed and liquid will
form. A decrease in pressure has caused liquid to form; this is the reverse of the behavior
one would expect, hence the name retrograde condensation. As the pressure continues
decreasing, more liquid will form until at some pressure the amount of liquid starts
decreasing. Eventually the dew point curve will be crossed again.
Dense Phase Region
It is common practice to refer to the area above the cricondenbar as the dense phase
region. In this region it possible to move from a temperature well below the critical
temperature to one well above it without any discernible phase change having taken
place. At the lower temperature the fluid would behave more like a liquid and at the
H-0806.35
4-5
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
higher temperature it would behave more like a vapor, but in between it would not exhibit
any of the traditional signs of a phase change.
4.2.2
Waxes
Naphthenes
Aromatics
Large molecules composed mainly of aromatic rings or carbon and hydrogen but
also can contain nitrogen, sulfur, oxygen, and metals
Water
Sulfur compounds
Nitrogen (N2 )
Helium
Metals
4.2.3
Vanadium, nickel
Mineral salts
H-0806.35
4-6
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The reservoir fluid type can be confirmed only by observation in the laboratory; however,
some rules of thumb can help identify the fluid type. Three properties that can be used
with the rules of thumb are the initial producing gas-oil ratio, the gravity of the stock tank
oil, and the color of the stock tank oil.
The behavior of a reservoir fluid during production is determined by the shape of its
phase diagram and the position of its critical point. Each of the five reservoir fluid types
can be described in terms of its phase diagram.
Black Oils
The phase diagram for a black oil is presented in Figure 4.2-3. Indicated on the phase
diagram is the critical point and quality lines. The vertical line in the figure indicates the
pressure reduction at constant temperature that occurs in the reservoir during production.
As the reservoir of a black oil is produced, the pressure will eventually drop below the
bubble point curve. Once below the bubble point, gas evolves from the oil and causes
some shrinkage of the oil.
Black oils are characterized as having initial gas-oil ratios (GORs) of 2000 SCF/STB or
less. The producing gas-oil ratio will increase during production when reservoir pressure
drops below the bubble point pressure. The stock tank oil will usually have a gravity
below 45API.
hydrocarbons.
The stock tank oil will be very dark due to the presence of heavy
Volatile Oils
The phase diagram for a typical volatile oil, Figure 4.2-4, is somewhat different from the
black-oil phase diagram. The temperature range covered by the two-phase region is
somewhat smaller, and the critical point is much lower than for a black oil and is
relatively close to the reservoir temperature (but still greater than the reservoir
temperature). The vertical line in the figure shows the reduction in reservoir pressure at
constant temperature during production. For a volatile oil, a small reduction in pressure
below the bubble point can cause a relatively large amount of gas to evolve.
The dividing line between black oils and volatile oils is somewhat arbitrary. Volatile oils
may be identified as having initial producing GORs between 2000 and 3300 SCF/STB.
The stock tank oil gravity is usually 40API or higher, and the stock tank oil is colored
(usually brown, orange, or green).
H-0806.35
4-7
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Retrograde Gases
The phase diagram of a retrograde gas, Figure 4.2-5, has a somewhat smaller temperature
range than that for oils, and the critical point is further down the left side of the phase
envelope. The changes are a result of retrograde gas containing fewer heavy
hydrocarbons than the oils. Additionally, the critical temperature is less than the
reservoir temperature, and the cricondentherm is greater than the reservoir temperature.
During initial production, the retrograde gas is single-phase gas in the reservoir. As the
reservoir pressure declines, the dew point is reached, and liquid begins to condense from
the gas and form a free liquid in the reservoir. This liquid will normally not flow and
cannot be produced.
The initial producing GORs for a retrograde gas ranges from 3300 SCF/STB on the lower
end to over 150,000 SCF/STB (the upper limit is not well defined). The producing GOR
will increase after the reservoir pressure drops below the dew point. Stock tank gravities
of the condensate are between 40 and 60API and increase as reservoir pressure drops
below the dew point. The stock tank liquid will be lightly colored to clear.
Wet Gases
With wet gases the entire phase envelope will be below the reservoir temperature as
illustrated in Figure 4.2-6. Wet gases contain predominately low molecular weight
molecules. A wet gases will remain as single phase gas in the reservoir throughout the
production life; however, the separator conditions do lie within the two-phase region.
Thus, somewhere in the production system, the dew point curve will be crossed and
liquid will condense from the gas.
Wet gases produce stock-tank liquids with gravities ranging from 40 to 60API; however,
the gravity of the liquid does not change during the production life. Wet gases have very
high GORs, typically more than 50,000 SCF/STB.
Dry Gases
Dry gases are primarily methane with some light intermediates. Figure 4.2-7 shows that
the two-phase regions is less than the reservoir conditions and the separator conditions.
Thus no liquid is formed in either the reservoir or the separator.
H-0806.35
4-8
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
4500
BLACK OIL
4000
PRESSURE PATH
IN RESERVOIR
3500
90
Pressure (psia)
3000
2
CRITICAL
POINT
80
2500
70
60
2000
50
1500
40
30
1000
20
10 % LIQUID
500
SEPARATOR
0
0
100
200
300
400
500
600
700
800
900
Temperature (F)
H-0806.35
4-9
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
4000
PRESSURE PATH
IN RESERVOIR
1
3500
VOLATILE OIL
CRITICAL
POINT
Pressure (psia)
3000
BUBBLE POINT
90
CURVE
80
70
2500
60
50
40
30
20
2000
10 % LIQUID
1500
1000
500
SEPARATOR
0
0
100
200
300
400
500
600
Temperature (F)
H-0806.35
4-10
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
5000
PRESSURE PATH
IN RESERVOIR
RETROGRADE GAS
Pressure (psia)
4000
CRITICAL
POINT
3000
60
50
40
30
20
10 % LIQUID
2000
3
SEPARATOR
1000
0
0
100
200
300
400
500
600
Temperature (F)
H-0806.35
4-11
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
3000
PRESSURE PATH
IN RESERVOIR
1
WET GAS
2500
Pressure (psia)
2000
CRITICAL
POINT
1500
20 % LIQUID
10
5
1000
2
SEPARATOR
500
0
-50
50
100
150
200
250
Temperature (F)
H-0806.35
4-12
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
3000
PRESSURE PATH
IN RESERVOIR
DRY GAS
2500
Pressure (psia)
2000
1500
1000
SEPARATOR
1% LIQUID
500
0
-100
-50
50
100
150
200
Temperature (F)
H-0806.35
4-13
1-Dec-00
4.3
4.3.1
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The black oil approach to the prediction of phase behavior ignores the fluid composition
and simply considers the mixture as consisting of a gas and liquid phase in which the gas
may be dissolved in the liquid. The basic assumption of a black oil model is that
increasing system pressure (and reducing temperature) cause more gas to dissolve in the
liquid phases, and, conversely, decreasing system pressure (and increasing temperature)
cause gas to evaporate from the liquid phase. It was previously noted that retrograde
condensation involves the conversion of gas to liquid on reducing pressure. This is
contrary to the fundamental assumption of the black oil model and so the black oil
approach is only valid for systems operating at conditions far removed from the
retrograde region.
In a typical liquid reservoir, the reservoir condition is well to the left of the critical point
and so the expansion process involves the continual evolution of gas, i.e. the operating
point moves steadily across the quality lines to a condition of ever decreasing liquid
content. This type of process would be adequately represented by a black oil model.
For the gas reservoir, the reservoir condition lies to the right of the critical point so that
on expansion, (reducing pressure) the operating point moves across the quality lines to a
condition of increasing liquid content, i.e. retrograde condensation. This process could
not be represented by a black oil model.
As a general guide a black oil model should be adequate for describing crude oil- gas
systems, while a compositional model is necessary to describe wet-gas, gas-condensate
and dense phase systems.
The black oil model employs certain concepts and nomenclature, which require
definition. These are discussed briefly below:
Producing Gas Oil Ratio (GOR)
This is the quantity of gas evolved when reservoir fluids are flashed to stock tank
conditions. The units are standard cubic feet of gas per stock tank barrel of oil
(SCF/STB) measured at 14.7 psia and 60F.
The GOR of a crude is obtained by experimental testing. However, the GOR will vary
depending on how many flash stages are employed to get down to stock tank conditions.
H-0806.35
4-14
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The normal convention is to calculate GOR from the sum of gas volumes evolved from a
multistage flash procedure (normally this involves 2 or 3 flash stages). This more closely
represents conditions in the field with the pressure and temperature conditions chosen for
the first stage flash approximating to conditions likely to be experienced in the first stage
separator in the field.
Solution Gas Oil Ratio (R s )
This is the quantity of gas dissolved in the oil at any temperature and pressure. It
represents the quantity of gas that would be evolved from the oil if its temperature and
pressure were altered to stock tank conditions, 14.7 psia and 60F. Hence, by definition
the Rs of stock tank oil is zero.
The Rs crude at its bubble point is equal to the producing GOR of the reservoir fluids.
The volume of free gas present at any pressure and temperature is the difference between
the GOR and the Rs. The volume of free gas is corrected for pressure, temperature and
compressibility to compute the actual in-situ volume of gas and hence superficial gas
velocity. Rs can be evaluated from standard correlations such as Glaso or Standing.
These correlations require as input the oil and gas gravity and the pressure and
temperature conditions.
Volume Formation Factor (B o )
The volume formation factor is the ratio of the volume occupied by oil at any pressure
and temperature to the volume occupied at stock tank conditions. The units are pipeline
barrels per stock tank barrel (BBL/STB). The volume formation factor of stock tank oil
is thus 1.0. Through use of Bo the volume flow rate and density of the liquid phase can
be calculated. Standard correlations are available to compute Bo . These require as input
the oil and gas density, the Rs of the liquid at the conditions of interest, and the pressure
and temperature.
Live Oil Viscosity
The viscosity of the oil in a two-phase pipeline depends on the stock tank oil viscosity
(dead oil viscosity), the solution gas oil ratio at the conditions of interest, and the pressure
and temperature. Correlations are available to compute the live oil viscosity.
The correlations available for Rs, Bo and live oil viscosity will yield approximate values
only and where laboratory or field data is available, these should be used to adjust and
H-0806.35
4-15
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
tune these correlations. The way in which the correlations are tuned will depend on the
quantity of field data available.
The minimum physical property information required to run a black oil model is:
sg i * Gi
i =1
Gi
i =1
where:
sgi = gravity of the ith separator stage off- gas
Gi = free gas GOR at the ith separator stage
n = number of stages in the separator train with the final stage at stock tank
conditions.
H-0806.35
4-16
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
applications is a minimum of production data: oil gravity, gas gravity, solution gas/oil
ratio and, if water is present, the watercut.
When to Use Black Oil Fluid Modeling
Black oil fluid modeling is appropriate for a wide range of applications and hydrocarbon
fluid systems. In general, the basic black oil correlations will provide reasonable
accuracy in most PVT fluid property evaluations over the range of pressures and
temperatures likely to be found in production or pipeline systems. However, care should
be taken when applying the "black oil approach" to highly volatile crude oils or
condensates where accurate modeling of the gaseous "light ends" is required. In this
case, the modeler needs to consider using compositional modeling techniques, which
describe the fluid as a multi-component system.
To increase the accuracy of the basic black oil correlations for modeling multiphase flow,
thermal- hydraulic simulators typically provide the facility to adjust salient values of a
number of the most important PVT fluid properties to match laboratory data.
Specifically, the following points can be calibrated:
Formation volume factor at pressure above the bubble point to account for oil
compressibility above bubble point
The above fluid properties are considered the single most important parameters affecting
the accuracy of multi-phase flow calculations. Calibration of these properties at the
bubble point and above can increase the accuracy of the correlations over all pressures
and temperatures.
This facility is typically optional, but the above calibrations will significantly improve the
accuracy of the predicted gas/liquid ratio, the flowing oil density and the oil volume
formation factor as a function of temperature and pressure. If the calibration data are
omitted, however, the thermal-hydraulic simulators will calibrate on the basis of oil and
gas gravity alone and thus, there will be a loss in accuracy. It should be noted that the
H-0806.35
4-17
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Black Oil calibration is only applicable to oil fluid types as it is not appropriate for a gas
fluid type.
4.4
Compositional Models
4.4.1
Equations of State
In a compositional model the predictions of gas and liquid physical properties are
performed through use of an equation of state, EOS. Any equation correlating pressure
(P), volume (V) and temperature (T) is known as an EOS. For an ideal gas the EOS is
simply:
PV = nRT
where:
n = number of moles of gas
R = Universal gas constant.
A gas is ideal if its molecules do not interact with each other and occupy no volume.
This is obviously not true, but the behavior of most real gases does not deviate drastically
from the behavior predicted by the ideal gas behavior. One way of writing an equation of
state for a real gas is to insert a correction factor into the ideal gas equation. This results
in:
PV = ZnRT
where the correction factor, Z, is known as the compressibility factor or z- factor. The
compressibility factor is the ratio of the volume actually occupied by a real gas at a given
pressure and temperature to the volume it would occupy at the same pressure and
temperature if it behaved like an ideal gas. The compressibility factor is not a constant.
It varies with changes in composition, pressure, and temperature.
To account for the non- ideality of most gas systems the ideal gas equation is modified to
include various correlating constants. The most commonly used equations of state used
in the oil and gas industry are called cubic equations of state because their mathematical
forms are cubic in terms of density or the z- factor. The two most popular equations of
state used in industry today are the Redlich-Kwong-Soave, the Peng-Robinson EOS, and
modifications of them.
H-0806.35
4-18
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
These cubic equations of state include terms to correct pressure for the forces of
attraction between the molecules. The actual pressure exerted by a real gas is less than
that of an ideal gas. Additionally, the cubic equations of state attempt to correct the
molar volume due to the volume occupied by the molecules.
The Peng-Robinson (PR) EOS, for example, is given by:
P =
RT
V b
a(T )
V (V + b) + b(V b )
b = i bi
i
k ij =
i j (l kij ) ai a j
and j.
ai a j =
temperature and represent the pressure contribution from the attractive forces.
The cubic equations of state can model liquids as well as gases and can be used to
calculate the vapor- liquid equilibria of hydrocarbon mixtures. The equation of state
allows a thermodynamically consistent method to evaluate the gas and liquid properties
when these two phases coexist.
The prediction of liquid densities was an area that ne eded improvement in original
development of the cubic equations of state. An empirical but effective way to improve
the accuracy of the liquid density predictions is to use the volume translation correction.
The volume translation is a linear correction of the predicted EOS volumes which does
not affect the equilibrium results from the original EOS. Therefore, this correction,
which is sometimes referred to as the Peneloux correction, is thermodynamically
consistent.
H-0806.35
4-19
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Another equation of state that is sometimes used in the oil and gas industry is the
Benedict-Weber-Rubin (BWR) equation and its derivative, the BWRS equation.
4.4.2
Viscosity
Viscosity, which is a transport property, cannot be evaluated from an EOS, but the EOS
provides compositional and property data that is needed in the viscosity models. Two
compositional methods to predict viscosity are commonly used: the LBC method (gas
and liquid) and the Pedersen method (gas and liquid). Preliminary testing has shown the
Pedersen method to be the most widely applicable and accurate for oil and gas viscosity
predictions. The Pedersen method is based on the corresponding state theory, as is the
LBC method.
Lower Alkanes
Predicted liquid viscosities using LBC and Pedersen methods have been compared to
experimental data for methane and octane as a function of both temperature and pressure
and for pentane as a function of temperature. For both methane and pentane the Pedersen
method predictions show close agreement with experimental data. For octane, the
Pedersen and LBC methods give comparable results. For the aromatic compound, ethyl
benzene, the Pedersen method is not as good as the LBC method.
Higher Alkanes
The results for higher alkanes eicosane and triacontane are mixed: the Pedersen method is
adequate for eicosane whereas the LBC method is slightly better than Pedersen for
triacontane. For triacontane both LBC and the Pedersen methods are inadequate.
However, in the majority of cases the higher hydrocarbons should be treated as petroleum
fractions rather than as single named components.
Petroleum Fractions
The LBC method describes viscosity as a function of the fluid critical parameters,
acentric factor and density. The LBC model is therefore very sensitive to both density
and the characterization of the petroleum fractions.
Water
The Pedersen method suffers the same drawback as the LBC method in that it is unable to
predict the temperature dependence of water, a polar molecule. To overcome this
problem, the Pedersen method has been modified especially for water so that it can
H-0806.35
4-20
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
accurately model the viscosity of water in the liquid phase. This was achieved by the
introduction of a temperature-dependent correction factor. However the prediction of the
viscosity of the gas phase is also affected, though in only a minor way.
Methanol
Neither the LBC nor the Pederson method can deal with polar components with the
Pederson method slightly worse than the LBC method. This is not surprising, as both
methods were developed for non-polar components and mixtures. The Pedersen method
works best with light alkanes and petroleum mixtures in the liquid phase. It performs as
well or better than the LBC method in nearly all situations.
The choice of the equation of state has a large effect on the viscosities predicted by both
methods. The LBC method is more sensitive to these equation of state effects than is the
Pedersen method.
4.5
Fluid Characterization
Petroleum reservoir fluids consist of thousands of different hydrocarbon molecules. The
diversity in chemical structure of the individual components increases with the carbon
number. In reality it is not practical to analyze for all of the components that may exist in
a reservoir fluid. Even if the separation and identification of each component present
were possible, the usefulness of such information would be limited. From a modeling
standpoint, it is desirable to keep the number of components small in using EOS to
minimize computation time requirements and round-off errors.
Standard composition analyses often stop at C7, C10, or C20. The gas chromatographic
analysis of pure hydrocarbon components up to C6 is routine. The physical and chemical
properties of these compounds (as required by an EOS) are accurately known. However,
compounds with higher carbon numbers are conventionally analyzed in terms of true
boiling fractions. The analysis is usually done in a gas chromatograph and provides the
mole fraction of all compounds that contain the same number of carbons in their
structure.
There are components that are too heavy and/or polar and are not volatile enough to be
separated by GC carbon number analysis. These components typically make up the
residue that is reported as the last carbon number component, and this residue consists of
all the components that have carbon numbers equal to or higher than the highest
H-0806.35
4-21
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
unseparable carbon number group. The residual group may be the C7+, C10+, C20+, or
C30+ fraction.
Because the components with carbon numbers C7 and higher are not separated as pure
compounds, their critical parameters are not known for use in EOS modeling. As a
result, a process is used to develop a set of pseudo-components to represent these
compounds and to determine the critical and other EOS parameters for these pseudocomponents. This process is referred to as the fluid or oil characterization process. An
EOS characterization refers to a list of hydrocarbon components and pseudo-components
and their critical properties and molecular weights, and it includes the binary interaction
parameters.
The fluid characterization procedure uses experimental data to assign equation of state
parameters to a set of pseudo-components. The experimental data often originates from
PVT experiments (e.g. constant mass expansion, constant volume depletion, differential
liberation, multistage flashes) of the reservoir fluid of interest. Viscosity data may also
be used. Because the characterization process will be using data for a specific reservoir
fluid, the resulting characterization will only be valid for that reservoir fluid. There are
no universal fluid characterizations.
The development of an EOS characterization proceeds through a series of steps:
All relevant experimental data is collected and reviewed. These data may include:
Constant mass expansion
Constant volume depletion
Differential liberation
Multistage flashes
Viscosity
Compositional analysis
H-0806.35
4-22
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Tuning EOS models to the experimental PVT data can be more of an art than a science,
and it requires the use of appropriate software programs. This is at least partially a result
of the EOS models being highly nonlinear and the number of adjustable parameters in the
regression being large. Additionally, there is no rigorous way to arrive at the global
minimum of such a highly nonlinear function. Special non-linear regression techniques
have been developed that allow adjusting the constants of the EOS and the critical
properties of the pseudo-components to tune the EOS predictions to PVT measurements.
There are limitations associated with fluid characterizations. The pseudo-components are
assumed to behave as single, lumped components in phase behavior, but in reality they do
not. Some of the pure components lump ed in a pseudo-component may not in reality
move from one phase to another as the pseudo-component does in the simulation of the
fluid. To overcome inaccuracies in the use of EOS to describe the phase behavior of
reservoir fluids, characterization procedures need to be followed to generate the most
appropriate set of pseudo-components and their relevant properties.
The EOS characterization may only be applicable to some of the processes the fluid may
undergo (e.g. reservoir depletion, flowline transport, facilities processing). These
processes may be those for which data were available and used in the development of the
characterization. Thus, the range of applicability of the EOS characterization depends on
the type of PVT data used and the pressure and temperature range of that data
4.6
H-0806.35
4-23
1-Dec-00
4.6.1
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Overview
Executive Summary
As field developments move to deeper water and subsea technology becomes more
widely used, paraffin and asphaltenes become more of a real problem than an annoyance.
Proper planning becomes critical and cannot be performed without data obtained from
representative fluid samples. CTR 901 was formed to address the special considerations
associated with collecting and handling fluid samples containing paraffins and
asphaltenes.
These guidelines were expanded somewhat beyo nd the basic goal of fluid sampling for
paraffin because it was recognized that in many instances the same sample would be used
for multiple reservoir fluid studies by a wide range of disciplines.
The following guidelines were developed with input from industry experts and with
vendor input. Issues related to sampling at surface facilities, sampling with downhole
flowstream samplers and sampling with downhole formation testers were addressed
individually.
Conclusions
In addition to the obvious concerns with obtaining a representative sample from the
reservoir, other problem areas must be understood and carefully addressed. First, all
equipment used in a sampling operation must be clean. Steam cleaning alone may not
remove previously deposited solids and these solids, which precipitate from one sample,
may dissolve in the next. Second, sample transfers are a major concern in the area of
sampling. In general, transfers performed on samples stabilized at reservoir conditions of
temperature and pressure should provide the greatest opportunity for representative
transfer. Response from the vendor community is that this is a realistic and attainable
goal. Consequently, vendor efforts have recently been directed toward the design and
testing of such a system. Ideally, proper planning and equipment selection can minimize
the number of transfers.
A major hindrance to getting samples to the lab exists in the area of availability of D.O.T.
approved transportation cylinders. While laboratories are increasing their capabilities to
analyze samples at reservoir pressure, the availability of suitable transportation cylinders
is lagging, especially above 10,000 psi. Vendors report that the cost and time associated
with obtaining D.O.T. approval for a specific cylinder design in the pressure ranges
required for deepwater development is prohibitive. While some vendors are pursuing this
H-0806.35
4-24
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
approval, none are currently known which can transport a sample at pressures higher than
10,000 psi.
Recommendations
4.6.2
Investigate all reasonable sampling options and carefully plan and document all
sampling operations. Coordinate planning efforts with all departments involved in
acquiring the sample or in the use of the data that will come from the sample.
Develop a prioritized analysis program for the sample detailing which analyses are
the primary purpose of obtaining the sample. Communicate with all vendors
involved in obtaining and analyzing the sample.
Minimize the number of transfers a sample will undergo. Perform transfers as near to
reservoir conditions of temperature and pressure as possible.
Do whatever possible onsite to verify that a satisfactory sample has been obtained
before concluding the sampling operation.
H-0806.35
4-25
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Studies/Chemical
Inhibitor
Stud ies
H-0806.35
4-26
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
6. Water analysis, e.g. chlorides, scale, corrosion (Reservoir Management and System
Design)
7. Crude Assay (Refinery Information, Product Value Determination)
Fluid types may or may not be known before sampling takes place. Certain sampling
methods can be problematic for specific fluid types. In addition to the anticipated phase,
information may also be available concerning contaminants like H2 S, CO2 , Sulfur, etc.
It is important to attempt to tabulate how much sample is needed to accomplish the goals
and objectives listed above. In addition to the quantity needed for a specific set of goals
and objectives, backup samples may be needed. A table is included in section 4.4.14
which may be of assistance in determining required sample volumes.
It is important to give prior thought to the equipment that will be needed or available for
a particular job. This applies not only to sampling equipment but also to any site transfer
and transportation equipment. Company policy and experience may limit choices in this
area. It should be verified that all necessary equipment is available and suitable for the
job. Among the things to check are:
Pressure and temperature ratings of all equipment. Verify with vendors that
pressurized tools can be heated to the desired temperatures for site transfer as well as
being rated for downhole conditions.
Verify that sample containers for transfer and storage meet the goals and objectives of
the job. Items to consider in the selection of sample containers include:
Whether atmospheric, low pressure (i.e. in the range of separator conditions) or
high pressure (i.e. in the range of reservoir conditions) will be required. Verify
that all cylinders will be pressure tested prior to use.
Whether special cylinders are required (e.g. for H2 S, Hg, etc.)
In all but rare instances D.O.T. certification of transportation is required. Not all
currently available equipment, especially in the higher pressure ranges, has been
approved for transportation in the United States. Verify with the vendor that all
necessary equipment has been D.O.T. certified (or exempted).
A variety of transfer and displacement mechanisms are available in sampling and
transportation equipment. Company policy and experience may limit the available
choices as well as safety concerns. The following list details the available transfer
H-0806.35
4-27
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Consider in advance any onsite transfer to transportation vessel needs. These may
include:
Having a sufficient quantity of equipment on hand including backup equipment in
case of problems.
Method and degree of heating. Coordination with sampling tool vendors will be
necessary to obtain a heating program which is acceptable to all parties.
Additional technology is needed in this area to provide heating methods that
address the safety concerns of the vendors related to doing transfers at higher
temperatures.
Solvents and other supplies for cleaning all equipment prior to and during the
sampling operation should be available along with proper disposal containers.
Make sure all vendor and field personnel are properly trained and understand the
importance of your sampling job. Sampling in existing developments is sometimes
performed by field personnel who may:
not be properly trained in sampling
not understand the importance of obtaining a representative sample and
maintaining it during transfer and analysis, and
not understand the importance of supplying proper documentation of the sampling
effort.
H-0806.35
4-28
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Make certain that the laboratories involved have the proper capabilities for the type of
sample you are providing. The additional cost of taking a pressurized sample and
transporting it under pressure to the lab is wasted if the lab must reduce the pressure
of the sample to perform the transfer or to analyze the sample.
H-0806.35
4-29
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
If possible avoid splitting sample until main goals are achieved. Splitting samples
and recombining them has inherent opportunities for sample alteration.
Cost and time limitation are always a consideration and may limit the type and
volume of sample taken.
Reservoir and well specific characteristics will impact your sampling efforts.
Following are some items to consider:
Wellbore: hole diameter, rugosity, deviation, size of casing and other well bore
equipment, drilling problems which have been encountered, etc.
Again, verify that all equipment is rated for the reservoir temperature and pressure
anticipated in this wellbore. Also, verify that all equipment is rated for any special
contaminants anticipated.
If possible make some prediction concerning the maximum drawdown that can be
achieved without taking reservoir fluid through a phase change. Often this will
not be possible.
Formation: Formation pressure, permeability, formation consolidation and grain
size.
Mud system: Mud system, mudcake and their associated filtrates and fines. In
some cases, critical sampling needs may dictate in advance that the mud system
meet certain criteria, e.g. it is extremely unlikely that an uncontaminated oil
sample can be acquired with formation testers if oil based mud is used in the
drilling of the well. In some cases, it has been reported that even after two weeks
of drill stem testing oil based mud contamination could still be detected in the
flow stream.
4.6.4
Determine in advance what will be needed in the "Final Sampling Report" and
communicate this to all relevant parties. Include specific requirements for
presentation of data and conclusions as well as for onsite documentation of the
sampling job.
Surface Sampling
Pre-job Preparation
Verify that the well is properly conditioned for sampling (See section 4.4.9).
Verify all equipment has been properly cleaned (See section 4.4.10).
H-0806.35
4-30
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Verify that sufficient sample containers of appropriate type are available and on site.
Verify that all sampling equipment is prepared for sampling.
Job Execution
Sampling points on the surface depend on the objective of sampling and tests to be
performed. Examples of sampling locations for various test objectives are as follows:
The wellhead or choke manifold may be the best sampling point when checking
(qualitatively) for the existence of paraffins and asphaltenes. This would typically be
the surface sampling point usually having the highest temperature and pressure with
the least likelihood of deposition having occurred. Care must be taken with high
pressure environments by using appropriate high pressure sampling cylinders. This
sampling point is also feasible for dead oil sampling.
The separator is the most suitable place to sample if the objective is to reconstruct the
reservoir fluid. This would be done for such tests as PVT, hydrates etc. Consider that
the test separator may contain contaminants from previous testing. Attempt to
properly size the separator to allow sufficient throughput to clean any residue left in
the separator. The primary stage separator should be the one used for sampling.
Sampling points on the separator include:
Siphon tube - A siphon tube is available on some separators which extends from
an external sampling valve down into the oil pan of the separator.
Oil dump - oil
Meter runs - oil or gas
Top valve - gas
Sight glass for oil or gas. This may not be preferred if it is cooler than the rest of
the separator.
H-0806.35
4-31
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Verify valves are plugged on arrival and for shipping. Sample containers should also be
checked for leaks prior to shipping. This can be done by checking for bubbles after
applying "Snoop", Soapy water, or by submersion in water.
Properly label all cylinders and document relevant details of the test. An example form
has been included in the section 4.4.13 which may be sent to the location of the sampling
job.
4.6.5
Location of the hydrocarbon sample with respect to the depth of the formation water
level to assure a representative sample.
H-0806.35
4-32
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Pump through modules or clean up chambers. The pump through modules allow
flowing an unlimited amount of sample through the tool to remove near probe
contamination. Clean up chambers allow flowing a limited volume of reservoir fluid
to auxiliary chambers.
Use fluid analyzers which can detect various differences in the fluid flowing into the
tool. Some currently available analyzers use either the resistivity or optical
characteristics of the fluids to make this differentiation.
Minimize the pressure drop while filling the sample chamber. Effectively this involves
the use of water cushions, throttling valves or chokes which may result in a longer
sampling period. Coordination with the drilling department will be necessary to arrive at
a mutually agreeable time period. Also, attempt to fill all void space within the tool with
water to prevent excessive drawdown at the instant the tool is opened.
It would be desirable to only sample one zone per run, even with a multi-sampler tool to
maximize the potential of taking an uncontaminated sample. These tools have portions of
the flow path which will be used for every sample taken. Sampling multiple zones in a
single run will cause some mixing of sample. The multi-sampler tools are better suited to
taking larger amounts of sample from a single zone.
Caution should be used to prevent pressure release during tool disassembly and sample
transfer at surface. This is a common sense statement but once pressure has accidentally
been released the damage has been done. Always assume you have a quality sample
during the transfer process even if downhole sensors or leakage at the surface suggest
otherwise.
Record any indication during disassembly of tool of downhole fluids: oil, gas, mud,
water, etc. This can be an early indication of whether hydrocarbons have been sampled or
whether only drilling fluid has been sampled.
Keep detailed documentation of sampling job. An example form has been included in the
appendix which may be sent to the location of the sampling job (See section 4.4.13).
H-0806.35
4-33
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Minimize number of transfers fluid will have to undergo. Every time sample is
transferred the likelihood of having an altered sample increases. The following
suggestions should be considered:
Consider sampling tools with transportable D.O.T. certified sample chambers.
Chambers which can be detached from the sampling tool and shipped to the lab
with the sample intact prevent onsite transfers.
If possible use transportation cylinders which can hold all of sample cylinder
volume so lab recombination will not be necessary. Subsampling into multiple
chambers means doing more onsite transfers and also more transfers when the
transportation cylinders arrive at the lab.
Agitate heated sample and return to single phase before transfer to promote
homogeneity.
Disassemble sample chamber and "swab" out all remaining oil and solids. Place these
solids and the swabbing cloth in a D.O.T. certified glass container for later analysis.
Report observations.
H-0806.35
4-34
1-Dec-00
4.6.6
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Chemically rinse the sample chamber until clean and place these samples along with
a virgin sample of the chemical used for rinsing in separate D.O.T. certified
containers (not plastic) for later analysis. Advance coordination with the vendor will
be necessary to identify appropriate cleaning solvents. For a more complete
discussion on cleaning see section 4.4.10.
Compositional gradients may result in the static fluid column from the pressure and
temperature gradient in that column.
Water may begin to settle in the bottom of the wellbore which may result in sampling
water.
The well may be sampled while flowing but issues such as the following must be
considered:
H-0806.35
4-35
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Drawdown across the perforations may cause gas to be liberated and the resulting
sample may be nonrepresentative.
At very low rates slugging may occur; again this may result in a nonrepresentative
fluid sample.
Minimize the pressure drop while filling the sample chamber to increase your chance of
sampling single phase fluid.
Caution should be used to prevent pressure release during tool disassembly and sample
transfer at surface. This will limit the usefulness of the sample and could prove very
costly in resampling.
Record any indication during disassembly of tool of downhole fluids: oil, gas, mud,
water, etc. This can be an early indication of whether the proper fluid has been sampled
(See also section 4.4.3).
Fluid should be compressed to maintain or obtain single phase condition during transfer.
Take backup surface samples if possible. This should be relatively inexpensive and may
prove invaluable if the bottomhole sample quality is questionable.
Keep detailed documentation of sampling job. An example form has been included in
section 4.4.13, which may be sent to the location of the sampling job.
4.6.7
Verify that opening pressure is the same as it was at the well site at the temperature at
which it was performed at the wellsite.
Repeat P-V check (i.e. bubble-point) that was done on site and at that temperature.
Stabilize temperature and pressure of the live fluid samples at reservoir conditions.
H-0806.35
4-36
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Technology Gaps
CTR 901 believes that additional R&D and implementation are needed in the following
areas. The items listed are either not currently available or don't exist in sufficient
quantities to meet projected Gulf of Mexico needs. In some cases the technology exists
but is not consistently implemented.
A sight glass in the transfer lines should be used during transfer to verify sample
quality onsite.
D.O.T. approved high pressure cylinders (> 10,000 psi.) with or without piston
displacement mechanisms.
Pressure compensated transportation cylinders - these are needed for situations where
asphaltenes are suspected.
Sample chambers that are transportable and DOT certified, preferably that can remain
at the lab for extended periods of time. These sample chambers would be part of the
sampling tool that could be removed and transported to the lab without having to
perform an onsite transfer. Ideally, they would remain at the lab until the priority tests
are completed and verified, possibly 60 days or so. Some vendors currently offer this
service. Unfortunately, not all vendors offe r this type of equipment and the equipment
available is limited in quantity and size and priced at a level that makes storage at the
lab during analysis very expensive.
H-0806.35
4-37
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Single phase samplers - These samplers use a nitrogen cushion to maintain reservoir
pressure on a sample as it is brought to the surface and cools. These are available in
the international market but only on a limited basis in the domestic market.
Downhole fluid analyzers that can accurately detect the difference between
hydrocarbons and all mud systems including oil based muds and synthetic oil muds.
Safe methods of heating sampling tools to 300F at the surface for transfers Currently, safety concerns with uneven heating has prompted some vendors to limit
the level to which they will allow their tools to be heated at the surface. Heating
methods acceptable to the vendors and customers should be feasible.
Improved transfer systems are needed which address the concerns in the previous
item. Also remote transfer capability is attractive from a safety standpoint.
Improved probe/reservoir interface in open- hole sampling tools. This is one of the
more common points of failure in formation tester samples.
Ability to truly control drawdown - Formation testers are needed which provide for a
reliable, predetermined drawdown. It is desirable to fill all void spaces in the tool and
chamber with a non-contaminating fluid. Additionally, the ability to variably
pressurize the pathways and chambers in the tool prior to and during sampling is
desirable.
Enhanced wellsite analytical capabilities are needed to verify samples before the rig
and sampling company leave the wellsite.
Variable rate downhole pump with ability to vent to annulus above top packer for
cased hole formation testers - This would permit large quantities of reservoir fluid to
be pumped away from sampling point to minimize contamination.
Improved agitation systems for transfers (balls, etc.) - Often these are not available.
These are needed to promote homogeneous sample transfers, especially when paraffin
deposition in the sampling tool is a concern.
H-0806.35
4-38
1-Dec-00
4.6.9
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Well Conditioning
Proper well conditioning is necessary for obtaining representative samples. The following
includes a general discussion of fluid phase behavior. An understanding of reservoir
fluids and phase behavior will assist in preparing an appropriate well conditioning plan.
Type Of Hydrocarbon Reservoirs
Hydrocarbon fluids fall into two main categories at reservoir conditions: liquids or gases.
Liquid hydrocarbons are referred to as bubble-point oils, and gases are refe rred to as
dew-point fluids.
Bubble-Point Reservoirs
Reservoir hydrocarbon fluid types are determined by the location of the point
representing the initial reservoir pressure and temperature with respect to the P-T diagram
of the fluid contained therein. If the reservoir temperature is below the critical
temperature and the reservoir pressure is at or above the bubble-point curve, the fluids are
characterized as bubble-point oils. Bubble-point oils range from black oils with gravity
and GOR generally below 400 API and 2000 SCF/bbl, respectively, to volatile oils which
generally exhibit higher gravity and GOR The literature is not in full agreement
concerning the criteria that characterizes the transition from black oils to volatile oils.
Gas Reservoirs
If the reservoir temperature is above the critical temperature and the reservoir pressure is
at or above the dew point curve, the fluids are characterized as dew-point gases. Dewpoint gases range from gas-condensates which release liquid condensate in the reservoir
below the dew-point (retrograde condensation), to wet gases which require reduction in
both P and T for any liquid to drop out. Gas-condensates and wet gases are characterized
by progressively higher gravities and GORs. Dry gases do not yield hydrocarbon liquid at
surface conditions.
Saturation Pressure
Bubble-Point Reservoirs
Bubble point systems are characterized by the coexistence of a liquid phase and an
infinitesimal amount of gas phase in equilibrium. The saturation or bubble-point pressure
(Pb ) is the fluid pressure in a system at its bubble-point. Pb is a function of system
composition and temperature. Oil reservoirs which exist above their bubble-point are
referred to as "undersaturated." Oil reservoirs which are associated with a gas cap are
H-0806.35
4-39
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
saturated at the gas-oil contact. The gas-cap may or may not exhibit retrograde
condensate behavior. In a saturated reservoir, the oil becomes progressively
undersaturated with depth below the gas-oil contact.
Dew-Point Reservoirs
Dew-Point systems are characterized by the coexistence of a gas phase and an
infinitesimal amount of liquid phase in equilibrium. The saturation or dew-point pressure
(PD ) is the fluid pressure in a system at its dew-point, and is a function of system
composition and temperature. Above the PD , the fluid exists as a single gaseous phase,
and exhibits retrograde condensation as the reservoir pressure falls below the original
dew-point pressure.
Prediction of Pb or PD
Knowledge of the type of hydrocarbon reservoir and its saturation pressure (bubble-point
or dew-point) is important for successful design and implementation of the sampling
operation. Estimates of saturation pressure can be obtained by using one or more of the
following leads:
MER testing
Published correlations
Well Conditioning
Effect of Drawdown
Drawdown is defined as the difference between static reservoir pressure (Pe) and
bottomhole flowing pressure (Pwf), or delta p = Pe - Pwf . As long as Pwf is greater than or
equal to Pb or PD , single phase fluid will flow into the wellbore. Such fluid would be
representative of the original reservoir fluid.
Bottomhole sampling methods require that the fluid pressure at sampler depth be above
the saturation pressure.
H-0806.35
4-40
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Reservoirs where the drawdown will result in diphasic flow at or below sampling depth,
are best sampled at the surface separator.
Dry Gas Reservoirs
Constant gas composition with pressure drop prevails in reservoirs, tubing and at surface.
Gas well conditioning beyond cleanup stage for sampling is not necessary.
Adequate reservoir gas sample can be obtained at wellhead upstream (or downstream) of
choke.
It is advisable to determine "dry gas" state by sampling/analysis at high and low well
rates, and upstream (U/S) and downstream (D/S) of choke.
Wet Gas Reservoirs
Constant gas composition with pressure drop prevails in reservoirs, and possibly in
tubing and at surface U/S of choke.
As with dry gas reservoirs above, determine gas state U/S and D/S of choke. Generally,
wellhead sampling U/S of choke is representative.
The only well conditioning required involves initial well clean- up and flow at sufficiently
low choke to ensure high enough wellhead pressure for wellhead sampling.
If gas is two-phase U/S of choke, must sample gas and liquid at separator after reaching
constant GOR. Procedure would be similar to sampling gas condensate reservoirs.
Gas-Condensate Reservoirs
Pe substantially greater than PD with low drawdown assuring monophasic flow into the
wellbore:
Very rich gas: Can be sampled with bottomhole or surface sampling. Additional
sample may be needed to obtain sufficient liquid due to shrinkage when bottom hole
samples are taken.
H-0806.35
4-41
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Medium to lean gas: surface sample to ensure adequate size sample dire to very high
shrinkage.
Pe slightly above or equal to PD - Diphasic flow into wellbore and possible condensate
drop-out in wellbore region. Well conditioning should be performed as indicated for
monophasic flow. Surface sampling early in life of reservoir is recommended.
Oil Reservoirs
Pe > Pb with low drawdown assuring monophasic flow into the wellbore:
Pe slightly above or equal to Pb with finite drawdown causing diphasic flow in the
wellbore:
Condition well as indicated above and obtain surface oil and gas samples early in the
life of the reservoir.
Stabilized THP
H-0806.35
4-42
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Constant differential and static pressure readings on the gas offtake chart
It is preferred that sampling takes place when ambient temperature is not fluctuating and
is below that of separator temperature.
4.6.10 Tool/Cylinder Cleaning
All equipment, samplers, pumps, cylinders, lines, valves and fittings should be
thoroughly cleaned and free of hydrocarbons or other contaminants. The reader should
recognize that there are numerous methods, ma terials and chemicals that are acceptable.
Below we have listed some that are currently used by some DeepStar companies.
Valve packing glands should be limited to teflon, graphite, viton or vespel (polymide,
graphite, fluorocarbon putty) or other noncontaminating substances.
Flush with warm (~160F) toluene or other suitable solvent. Initially soak for 2
hours, agitate if feasible and drain. Continue process until effluent is visibly clear or
the refractive index as measured by light refractometer agrees with that for pure
toluene.
4-43
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 4-1: Example of apparatus for sampling wells cutting significant amounts of water.
H-0806.35
Raise pressure in bottle 1 by 1000-2000 psi above the estimated Pb. Agitate contents
during pressuring. A positive displacement pump is used to raise the pressure
utilizing and immiscible fluid for non-piston type containers, or any convenient fluid
for piston type containers.
4-44
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Heat bottle 1 to about 1800 F while maintaining the pressure above Pb.
Allow the bottle to stand in a vertical position overnight. This allows all water to
separate out at container bottom.
Connect bottle 2 as shown in the diagram incorporating a two way valve and about 20
ft. of line loop downstream of the valve. Evacuate line and valve.
Commence sample transfer from bottle 1 to bottle 2 ensuring Pressure is safely above
Pb. Initially fill valve and line loop with reservoir fluid and pressure up to transfer
pressure.
Sample the fluid stream (using the two-way valve between the bottles) every 20 ml.
The sample will consist of a drop or two only to test whethe r the line contains any
water. At the first sign of water in the transfer line, shut the two-way valve followed
by the top valve on bottle 2. Any overlooked water in the transfer line will be caught
in the loop.
H-0806.35
4-45
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Type of
Analysis
On-site
Description
Purpose
Minimum Qty
(cc)
Preferred Qty
(cc)
Flashed
or Dead
Flashed
or Dead
Live
Live
Other
Samples
Minimum Qty
(cc)
Visual Observation
verification
of
downhole
sample
Opening pressure
verification
Fingerprint for
contamination
10
250
250
To determine the
ability of production
to flow at various
temperatures and
pressures
100
200
To know fluid
properties at various
temperatures and
pressures
100
500
Compositional
analysis & GOR
Specific gravities
& API
Molecular weight
PVT
Constant-Mass
Expansion
Bubble point
Liberation of gas
Specific gravity
Compressibility
Viscosity
@ reservoir
conditions
Separator flash
Water
content in
live sample
ASTM (BS&W) D-96-88
container
Cloud
point
methodsSelect at
least one
H-0806.35
Microscopy
Fundamental test to
determine oil in
place and measure
further fluid
properties
10
production solids
50
Determine onset
temperature for wax
precipitation
Differential Scanning
Calorimetry (DSC)
4-46
10
10
10
10
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
STANDARD ANALYSES
Type of
Analysis
Total
Sulfur
Nickel &
Vanadium
Total Acid
Number
Pour Point
Description
Minimum Qty
Preferred Qty
Other
(cc)
(cc)
Samples
Flashed
or Dead
Purpose
Live
Flashed
or Dead
Light Transmittance
200
200
Cold Finger
200
200
200
200
20
20
20
20
10
10
50
160
Determine value of
crude and limitations
for transportation
A
high
Live
Minimum Qty
(cc)
content
devalues crude. A
contaminant for the
refinery
Determine value of
crude,
affects
refinery,
transportation,
potential solids
&
Determine
temperature that a
ASTM D97-93 (IP 1567/86) crude will gel and no
longer flow in static
conditions
High Temp
Gas
Chromatography
(HTGC)
Provides information
for wax deposition
10
20
%
ASTM D-3279-90 (IP
Asphaltene 143/84)
Screening value to
estimate potential for
operational
problems.
Percent
asphaltene
also
affects crude value.
15
15
Elemental
Analysis
(C, H, N,
O, S)
Geochemistry? Input
data for value of
crude
20
20
H-0806.35
ASTM D-5291
4-47
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
STANDARD ANALYSES
Type of
Analysis
Description
Wax
Content
UOP 46-85
PNA or
SARA
Liquid Chromatograph
analysis
Purpose
Screening value to
estimate relative wax
problems. Low value
does not mean a wax
problem will not
occur
General information
of crude composition
H-0806.35
Minimum Qty
Preferred Qty
Other
(cc)
(cc)
Samples
Flashed
or Dead
Flashed
or Dead
50
50
10
100
825
4-48
Live
460
1035
Live
Minimum Qty
(cc)
960
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Table 4-2: Volume Requirements for Detailed Solids Formation Studies and Other
Analyses
Type of Analysis
Description
Purpose
Flashed
or Dead
Flashed
or Dead
Live
Live
Other
Samples
Minimum
Qty (cc)
Paraffin
Rate of
deposition
Yield
stress (high
pour point
Determine severity of
wax problem Restart
pressure requirements
Determine severity of
problem
crudes)
Asphaltenes
Stability tests
Experimentally
Hydrate
Formation
conditions
measure
formation
conditions. Especially
for liquid dominated
systems.
Hydrate
formation potential
can be modeled using
25
50
PVT
data
and
reservoir properties.
OTHER ANALYSES
Water
Compositional
Determine potential
analysis
(API
for scale, corrosion,
RP45 rev in '96).
and
formation
Organic
acid
concerns
content
150
Drilling fluid
Compositional
analysis
100
Determine extent of
contamination
inorganic
composition
Solids
H-0806.35
Organic/
chromat
ographic
analysis
If
available
production
4-49
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Compositional
analysis
Identify extent of
contamination
and
100
sample integrity
Identify extent of
contamination
and
sample integrity
100
H-0806.35
4-50
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
4-51
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
4-52
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
inhibitor in the field. Small scale laboratory testing will typically require
approximately 50 mls per test.
4. For large scale flow loop tests, much larger volumes are required. If you are
considering a large scale test, contact the flow loop operator to identify the exact
sample requirements.
5. For gas reservoirs, it may be possible to use a synthetic gas sample based upon the
compositional analysis from the PVT report for the hydrate measurements. For black
oil or condensate systems, actual samples should be used. Standard sampling
techniques used for obtaining samples for PVT measurements are sufficient. Please
see the DeepStar Fluid Sampling Guidelines for further recommendations.
Scale Samples
Samples from all potential water sources (both formation and injection waters) are
necessary to allow identification of scaling tendencies. Ideally, 150 mls of each water is
required. Some suggested procedures are:
1. Obtain water samples before any mixing of different waters or before any scale
precipitation occurs. This may require downhole samples.
2. Consider the reservoir implications in identifying where and what to sample. Which
waters will be mixing, where?
3. For separator samples, collect the samples in glass or plastic bottles filled to
overflowing, capped and sealed with tape.
4. It is recommended that the pH, carbonate, bicarbonate and sulfide concentrations be
measured on-site, as these components will change over time.
5. Ideally two identical samples should be collected at the same time. One should be
sampled as described above. The other should be acidified to a pH <2 in a glass or
plastic bottle filled to the top and taped shut. This preserves samples against solids
forming so an accurate analysis of dissolved cations can be obtained at the lab.
A source for additional water sampling guidelines is the ASTM standard D-3370.
4.7
Information Sources
Barrufet, M. A., A Brief Introduction to Equations of State For Petroleum Engineering
Applications, Harts Petroleum Engineer International, March 1998.
H-0806.35
4-53
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Chew, J. & Connally, C.A. "A Viscosity Correlation for Gas Saturated Crude Oils"
Trans. AIME, 216, p23, 1959.
Christensen, P.L., Regression to Experimental PVT Data, J. Can. Pet. Tech., Vol. 38,
No. 13, 1999.
Glaso, O. "Generalized Pressure-Volume- Temperature Correlations" J. Petrol. Technol.,
p785, May 1980.
Lasater, J.A. " Bubble Point Pressure Correlation", Trans. AIME, 213, p379, 1958.
Leontaritis, K. J., PARA-Based Reservoir Oil Characterizations, SPE 37252.
McCain, W. D., The Properties of Petroleum Fluids, PennWell Publishing Co., Tulsa,
OK, 1990.
Pedersen, K. S., Fredenslund, A., and Thomassen, P., Properties of Oils and Gas
Consensate Mixtures; Gulf Publishing Co., Houston, Texas, 1989.
Peng, D. Y. and Robinson, D. B., A New Two-Constant Equation of State, Ind. Eng.
Chem. Fundam., Vol. 15, pp. 59-64.
Soave, G., Equilibrium Constants from a Modified Redlich-Kwong Equation of State,
Chem. Eng. Sci., Vol. 27, pp. 1197-1203, 1972.
Standing, M.B. "A Pressure-Volume-Temperature Correlation for Mixtures of California
OIls and Gases," Drill. Prod. Practice, API, p247, 1947.
Vasquez, M. and Beggs, H. D., "Correlations for Fluid Physical property Prediction,"
SPE 6719, J. Petrol. Tech., p968, 1980.
H-0806.35
4-54
1-Dec-00
5.0
MULTIPHASE FLOW
5.1
Introduction
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The simultaneous flow of gas and liquid through pipes, often referred to as
multiphase flow, occurs in almost every aspect of the oil industry. Multiphase
flow is present in wellbore tubing, gathering system flowlines, and processing
facilities and has become increasingly important in recent years due to the
development of marginal fields and deepwater prospects. In many cases, the
feasibility of these prospects/fields hinges on cost and operation of the pipelines
and the associated equipment. Multiphase flow characteristics can have a
profound effect on both the design and operation of these systems. For example,
the type of flow regime (described in Section 5.2) can effect the pressure drop
along a pipeline. The amount of the pressure drop can affect either pipe size
and/or pumping requirements depending on the scenario.
Multiphase flow in pipes has been studied for more than 50 years, with significant
improvements in the state of the art during the past 15 years. The best available
methods can predict the type of multiphase flow characteristics much more
accurately than those available only a few years ago. The designer, however, has
to know which methods to use to get the best results.
The objective of this section is to present the basic principals of multiphase flow
and illustrate the current methods available to predict multiphase flow regimes,
pressure drop and liquid holdup. This chapter is arranged in the following order.
Flow regimes are discussed and described in Section 5.2. Section 5.3 describes
the flow models used to predict the flow regimes. The calculation of pressure
drop and liquid holdup based on the flow regime is discussed in Sections 5.4 and
5.5. Section 5.6 illustrates the effects of three phases.
5.2
Flow Regimes
In multiphase flow, the gas and liquid within the pipe are distributed in several
fundamentally different flow patterns or flow regimes, depending primarily on the
gas and liquid velocities and the angle of inclination. In general, flow regimes are
split into two major categories based on geometry: horizontal and vertical. The
term "horizontal" is used to denote a pipe that is inclined in a range between plus
or minus 10 degrees. The term "vertical" denotes upward inclined pipes with
angles from 10 to 90 degrees from horizontal.
H-0806.35
5-1
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Observers have labeled these flow regimes with a variety of names. Over 100
different names for the various regimes and sub-regimes have been used in
literature. In this guide, only four flow regime names will be used: slug flow,
stratified flow, annular flow, and dispersed bubble flow. Figure 5.2-1 shows the
flow regimes for near horizontal flow, and Figure 5.2-2 shows the flow regimes
for vertical upward flow. Descriptions of the flow regimes are as follows:
5.2.1
Stratified Flow
Stratified Flow generally occurs at low flow rates in near horizontal pipes. The
liquid and gas separate by gravity, causing the liquid to flow on the bottom of the
pipe while the gas flows above it. At low gas velocities, the liquid sur face is
smooth. At higher gas velocities, the liquid surface becomes wavy. Some liquid
may flow in the form of liquid droplets suspended in the gas phase. Stratified
flow only exists for certain angles of inclination. It does not exist in pipes that
have upward inclinations of greater than about one degree. Most downwardly
inclined pipes are in stratified flow, and many large diameter horizontal pipes are
in stratified flow. This flow regime is also referred to as stratified smooth,
stratified wavy, and wavy flow by various investigators.
5.2.2
Annular Flow
Annular flow occurs at high rates in gas dominated systems. In annular flow part
of the liquid flows as a film around the circumference of the pipe. The gas and
remainder of the liquid (in the form of entrained droplets) flow in the center of the
pipe. The liquid film thickness is fairly constant for vertical flow, but it is usually
asymmetric for horizontal flow due to gravity. As velocities increase, the fraction
of liquid entrained increases and the liquid film thickness decreases. Annular
flow exists for all angles of inclinations. Most gas-dominated pipes in highpressure vertical flow are in annular flow. This flow regime is also referred to as
annular- mist or mist flow by many investigators.
5.2.3
H-0806.35
5-2
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Various investigators have also referred to this flow regime as froth or bubble
flow.
5.2.4
Slug Flow
Slug flow for horizontal flow usually occurs at moderate gas and liquid velocities.
Slugs form due to high vapor shear causing the development of waves on the
surface of the liquid to grow to a sufficient height to completely bridge the pipe.
When this happens, alternating slugs of liquid and gas bubbles will flow through
the pipeline. This flow regime can be thought of as an unsteady, alternating
combination of dispersed bubble flow (liquid slug) and stratified flow (gas
bubble). The slugs can cause vibration problems, increased corrosion, and
downstream equipment problems due to its unsteady behavior.
Slug flow also occurs in near vertical flow, but the mechanism for slug initiation
is different. The flow consists of a string of slugs and bullet-shaped bubbles
(called Taylor bubbles) flowing through the pipe alternately. The flow can be
thought of as a combination of dispersed bubble flow (slug) and annular flow
(Taylor bubble). The slugs in vertical flow are generally much smaller than those
in near horizontal flow. Slug flow is the most prevalent flow regime in low
pressure, small diameter systems. In field scale pipelines, slug flow usually
occurs in upwardly inclined sections of the line. It occurs for all angles of
inclinatio n. Investigators have used many terms to describe parts of this flow
regime. Among them are intermittent flow; plug flow; pseudo-slug flow, and
churn flow.
H-0806.35
5-3
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
5-4
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
5-5
1-Dec-00
5.3
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
5.3.1
H-0806.35
5-6
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
regime. Correlations for a few key parameters are required to solve the equation
set. Mechanistic models extrapolate to field conditions much better than the
multiphase flow correlations because the mechanistic models account for the
effects of all the major variables.
Several mechanistic models have been developed in the past few years. Tulsa
University has developed models for near vertical flow (Ansari) and a general
model covering all inclinations (Xiao). The physics in these models are good, but
the correlations built into them are based on small diameter, low-pressure data.
The OLGA-S mechanistic model is based on a wider range of data (diameters
from 1 to 8 inches and pressures from atmospheric to 1400 psi), and it is generally
believed to extrapolate best to field conditions.
On occasion, the conditions for a simulation may cause otherwise good
multiphase flow methods to give erroneous results. It is always a good idea to
check the results by use of another method to ensure that the answers are
reasonable.
The mechanistic methods simultaneously and consistently solve for both pressure
drop and holdup and as a result, can predict trends more accurately over a wider
range of variables (pressure, temperature, CGR, pipeline length and diameter)
than correlation based methods. A general failing of mechanistic models is their
poor performance for slugging and other intermittent flow regimes. If available,
comparison with field data or the results of suitable correlation methods is usually
advisable.
5.3.2
H-0806.35
5-7
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The transitions between the flow regimes are frequently depicted in a flow regime
map, such as those shown in Figures 5.3-1 and 5.3-2. The flow regime map
typically has the superficial gas velocity (Vsg) on the X-axis and the superficial
liquid velocity (Vsl) on the Y-axis. The flow regime map is only valid for a single
point in the pipeline. As the angle of inclination, pressure and temperature
change with position in the pipeline, the flow regime map also changes.
Experimental studies of flow regime transitions have shown that each of the flow
regime boundaries reacts differently to changes in the system variables. Table
5.3-1 shows the sensitivity of the transitions to changes in the major system
variables.
Table 5.3-1: Sensitivity of Flow Regime Transitions to System Variables
Transition
Variable
Slug to
Dispersed
Bubble
Slug to
Annular
Slug to
Stratified
Stratified to
Annular
Angle of
Inclination
Small Effect
Moderate
Effect
Strong Effect
Strong Effect
Gas Density
Small Effect
Strong Effect
Strong Effect
Strong Effect
Pipeline
Diameter
Small Effect
Small Effect
Strong Effect
Moderate
Effect
Liquid Physical
Properties
Moderate
Effect
Small Effect
Moderate Effect
Moderate
Effect
The designer needs to carefully choose the method that will work best for the set
of conditions. The best methods are discussed in the remainder of this section.
H-0806.35
5-8
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
5-9
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
5-10
1-Dec-00
5.3.3
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
5-11
1-Dec-00
5.4
5.4.1
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The pressure drop along a pipeline is one of the most important factors relating to
its operation. Accurate calculation of pressure drop is essential for the proper
sizing of the pipeline. An under-sized pipe will suffer excessive pressure drop
due to high fluid velocities, which can limit its throughput. An over-sized pipe
can suffer from excessive holdup, which can lead to slugging and may require a
high pigging frequency.
The design of a multiphase flow line requires knowledge of the physical
properties of the fluid, its phase equilibrium, the flow regimes which exist within
it and the profile of the route which it will follow. The through- life pressure drop
will indicate whether compression will be required at some point, and at what
point in the life of the field. This can seriously affect the economics of the field.
Unfortunately, pressure drop is one of the most difficult pipeline parameters to
calculate accurately.
In single-phase flow, flow resistance in a pipe is due primarily to pipe friction and
elevation differences. In multiphase flow systems, additional complexity arises
because frictional energy is not only dissipated at the pipe wall, but also at the
interface(s) between liquid and gas phases, usually resulting in slip between the
phases. In addition, whereas for single-phase gas flow, the pressure drop will not
be as dependent on the pipe elevation, in multiphase flow, it will be. In
multiphase flow, the overall rise height of the pipeline, along with local gradients
must be accounted for in the design.
The overall pressure gradient is composed of three additive elements:
H-0806.35
5-12
1-Dec-00
5.4.2
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Frictional Losses
In multiphase flow, frictional losses occur by two mechanisms: friction between
the gas or liquid and the pipe wall, and frictional losses at the interface between
the gas and liquid. The friction calculations, therefore, are highly dependent on
the flow regime, since the distribution of liquid and gas in the pipe changes
markedly for each regime.
In stratified flow, there is wall friction between the gas and the pipe wall at the
top of the pipe, and wall friction between the liquid and the wall at the bottom of
the pipe. There is also friction between the gas and liquid at the gas-liquid
interface. The interfacial friction can be similar in magnitude to the wall friction
if the interface is smooth, or it can be considerably higher if waves are present.
In annular flow, there is friction between the liquid film and the wall. There is
also considerable interfacial friction between the gas in the core of the pipe and
the liquid film. The interfacial friction is usually the larger component.
In dispersed bubble flow, friction occurs between the liquid and the wall. There is
negligible interfacial friction between the gas and liquid.
Slug flow has several frictional components. In the slug, the friction losses are
caused by the friction between the liquid and the pipe wall. In the gas bubble, the
frictional components are the same as in stratified flow, namely gas and liquid
friction with the pipe walls and interfacial friction between the gas and liquid.
The friction loss in the slug is usually much higher than the losses in the bubble.
5.4.3
Elevation Losses
Elevational losses may be the major pressure loss component in vertical flow and
flow through hilly terrain. The calculation of elevational losses is governed by
the following equation:
g sin
dp
= mix
dx elev
144g c
where:
(dp/dx)elev = Pressure gradient due to elevation, psi/ft
mix
H-0806.35
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
= Liquid Holdup
= Angle of inclination
gc
Acceleration Losses
Although acceleration losses are present for all flow regimes, they are only
significant for two flow regimes: annular flow and slug flow. The mechanisms
for the losses in these two flow regimes are very different and will be discussed
separately.
In single-phase flow, acceleration losses can be calculated from Bernoulli's
equation. Acceleration losses represent the change in kinetic energy as the fluid
flows down the pipe. The expression for acceleration gradient is:
V dV
dp
dx accel 144g c dx
where:
= Density, lb m/ft3
V = Velocity, ft/sec
For multiphase flow, the same type of relationship holds except that it refers to
the flow of the mixed phase fluid. Most methods assume a no-slip mixture and
use the no-slip mixture density (ns) and the mixture velocity (Vm) in the
calculation of acceleration losses.
The kinetic energy acceleration losses are small for most oil industry applications.
The main exception is high velocity flow through low-pressure piping. Flare
systems would be an example of piping that has high acceleration losses.
Acceleration may account for 30 to 50 percent of the overall pressure loss in such
lines. For a typical high pressure gathering system line, acceleration is usually
less than 1 percent of the total drop and is frequently ignored.
H-0806.35
5-14
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
5-15
1-Dec-00
5.4.6
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Pipewall Roughness
The relative roughness used here is defined as the absolute pipe roughness (which
can be thought of as the surface-to-peak height of undulations on the metal
surface) divided by the pipe diameter. This is then usually applied to a Moody
friction chart to find the friction factor. The Moody chart gives the Fanning
friction factor as a function of Reynolds number and relative roughness.
A new steel pipe is usually reckoned to have an absolute roughness of about 50
micron. In a 24- inch pipe this converts to a relative roughness of about .00008.
An aged, corroded pipe will have a much rougher sur face and a corresponding
higher relative roughness value. There can be a factor of about two between the
pressure drop that can be expected between the new and the aged pipe. Since
flow rate is roughly proportional to the square root of pressure drop, this implies
that the new smooth pipe could have a capacity 40 percent greater than the very
rough pipe.
It is possible to generate a smoother surface than 50 microns by the use of
internally lining or coating the pipe. A roughness (smoothness?) of 10 microns is
feasible by this means, converting to a relative roughness of about 0.000016. The
Moody chart shows that at the maximum Reynolds number this produces a
friction factor of about 0.00115. The increase in capacity that might be expected
by coating or lining the pipe is therefore 6 to 7 percent.
5.4.7
H-0806.35
5-16
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
5.5.1
H-0806.35
5-17
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
No-Slip Holdup
The no-slip holdup is calculated by assuming homogeneous flow throughout the
length of the pipeline. If there were no slip between the gas and liquid phases,
both phases would move through the pipe at the mixture velocity. The liquid
would occupy the volume fraction equivalent to the ratio of the liquid volumetric
flow rate to the total volumetric flow rate. In multiphase flow terminology, this
equates to the liquid holdup being equal to the ratio between the superficial liquid
velocity and the mixture velocity:
Hl,ns
where:
Vsl = Superficial Liquid Velocity (actual ft3 /sec of liquid / pipe crosssectional area)
Vm = Mixture Velocity
In practice, no-slip conditions can only be achieved at extremely high gas
velocities where the pressure drops are uneconomic and there is a risk of erosion.
The no-slip holdup thus represents the minimum holdup possible in a multiphase
pipeline. At typical pipeline operating conditions of 4C and 70 barg, the no-slip
holdup can be estimated as 0.045 percent of the pipeline volume for each 1
BBL/MMscf of CGR.
H-0806.35
5-18
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Under most conditions, the liquid phase, which is more dense and viscous, moves
more slowly than the gas. When this occurs, the liquid holdup (Hl) is greater than
the no-slip holdup. Under these conditions, the actual gas velocity is greater than
the mixture velocity, and the actual liquid velocity is smaller than the mixture
velocity. The expressions for the actual gas velocity (Ug ) and the actual liquid
velocity (Ul) are:
Ug =
Ul =
5.5.3
Vsg
1- H l
Vsl
Hl
Annular
Flow
Stratified
Flow
Dispersed
Bubble Flow
Superficial Gas
Velocity
Strong
Strong
Strong
Strong
Superficial
Liquid Velocity
Strong
Strong
Strong
Strong
Gas Density
Moderate
Strong
Strong
Strong
Pipeline
Diameter
Moderate
Weak
Weak
Weak
Angle of
Inclination
Moderate
Weak
Very Strong
None
Liquid
Properties
Moderate
Moderate
Moderate
Weak
H-0806.35
5-19
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
As seen in Table 5.5-1, the influence of the major variables on the holdup is very
different for each of the flow regimes. As a result, it is impossible to develop a
general holdup correlation that will apply to all the flow regimes. Unfortunately,
almost all of the commonly used ho ldup models available in commercial software
try to do this. They work poorly over much of the operating range as a result.
The only way to accurately predict liquid holdup is to use mechanistic models for
each of the flow regimes.
5.5.4
5.5.5
Mechanistic Models
Mechanistic models calculate liquid holdup as part of their overall solution to the
flow condition in each section of a pipeline. These models work at a more
fundamental level than the correlation methods as discussed in the prediction of
flow regimes.
5.5.6
H-0806.35
5-20
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
5-21
1-Dec-00
5.7
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Information Sources
Agrawal, S.S., Gregory, G.A. & Govier, G.W. " An Analysis of Horizontal Stratified
Two Phase Flow in Pipes" Can. J. Chem. Engng. p280-287, 51, June 1973.
Baker, A. & Gravestock, N. "New Correlations for Predicting Press Loss and Holdup in
Gas/Condensate Pipelines" BHRA 3rd Int. Conf. on Multiphase Flow, The Hague
18-20th May 1987.
Baker, A., Nielsen, K. & Gabb, A. "Field Data Test New Holdup, Pressure- loss
Calculations for gas, Condensate Pipelines" O&G J. p78-86, March 21th, 1988.
Baker, A., Nielsen, K. & Gabb, A. "Holdup, Pressure- loss Calculations Confirmed"
O&G J. p44-50, March 28th, 1988.
Baker, A., Nielsen, K. & Gabb, A. "Pressure Loss, Liquid-holdup Calculations
Developed" O&G J. p55-59, March 14th, 1988.
Beggs, H.D. & Brill, J.P. "A Study of Two Phase Flow in Inclined Pipes"
Trans.Pet.Soc.AIME, p607, 256, 1973.
Beggs, H.D. and Brill, J.P., A Study of Two-phase Flow in Inclined Pipes, J. Pet. Tech.,
June 1967, P. 815.
Cunliffe, R.S. "Condensate Flow in Wet- gas Lines can be Predicted" O&G J. p100-107,
30th October. 1978
Eaton, B.A., Andrews, D.E., Knowles, C.R., Silberberg, I.H. & Brown, K.E. " The
Prediction of Flow Patterns, Liquid Hold- up and Pressure Losses Occurring during
Continuous Two-phase Flow in Horizontal Pipelines" J. Pet. Tech., p815-828, June
1967.
Gregory, G.A. "Comparison of Methods for the Prediction of Liquid Holdup for Upward
Gas- liquid Flow in Inclined Pipes" Can. J. Chem. Engng. p384-388, 53, August 1975.
Gregory, G.A. "Multiphase Flow in Pipes - Notes for a Professional Development
Course Neotechnology Consultants Ltd., Calgary 1991.
Gregory, G.A. & Fogarasi, M. "A Critical evaluation of Multiphase Gas- liquid Pipeline
Calculation Methods" BHRA 2nd Int. Conf. on Multiphase Flow, London, 19-21st June
H-0806.35
5-22
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Calculation Methods" BHRA 2nd Int. Conf. on Multiphase Flow, London, 19-21st June
1985.
Hansen, T.E. "The Northeast Frigg Full Scale Multiphase Flow Test" BHRA Multiphase
Production Conference, Cannes, 19-21st June, 1991.
Hughmark, G.A. "Hold-up in Gas- liquid Flow" Chem. Engng. Prog. P62, 58, 1962.
Lockhart, R.W. & Martinelli, R.C. " Proposed Correlation of Data for Isothermal
Two-phase, Two Compone nt Flow in Pipes" Chem. Engng. Prog. p45, 39, January
1949.
Mandhane, J.M., Gregory, G.A. & Aziz, K. "A Flow Pattern Map for Gas- liquid Flow in
Horizontal Pipes" Int. J. Multiphase Flow, p537, 1, 1974.
Mandhane, J.M., Gregory, G.A. & Aziz, K. "Critical Evaluation of Friction
Pressure-drop Prediction Methods for Gas- liquid Flow in Horizontal Pipes" J. Pet. Tech.
p1348-1358, October 1977.
McAllister, J.S., Hydrocarbons GB Ltd. memorandum and attached technical note dated
15th April 1988.
Minami, K. & Brill, J.P. "Liquid Hold-up in Wet Gas Pipelines" SPE 14535, SPE
Production Engineering, p36-44, February 1987.
Molyneux, P.D. "The Measurement of Multiphase Flow in a Pipeline Dip and
Comparison of the Results with PLAC Predictions" BG R&T Internal Report GRC R
1360, April 1996.
Mukherjee H. & Brill, J.P. "Liquid Holdup Correlations for Inclined Two-phase Flow
J. Pet. Tech., p1003-1008, May 1983.
Mukherjee, H. and Brill, J.P., Pressure Drop Correlation for Inclined Two-phase Flow,
J. of Energy Resources Te ch., Vol. 107, P.549, 1985.
Oliemans, R.V.A. "Modeling of Gas-condensate Flow in Horizontal and Inclined Pipes"
Proc. ASME, Pipeline Engng. Symp., ETCE, p73, Dallas, February 1987.
Oliemans, R.V.A., Two-phase Flow in Gas Transmission Pipelines, ASME paper
76-Pet-25, Sep. 1976.
H-0806.35
5-23
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
5-24
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
6.0
6.1
Introduction
The majority of pipelines are sized by use of three primary design criteria:
available pressure drop; allowable velocities; and flow regime Line sizing is
usually performed by use of steady state simulators, which assume that the
pressures, flowrates, temperatures, and liquid holdup in the pipeline are constant
with time. This assumption is rarely true in practice, but line sizes calculated from
the steady state models are usually adequate. Thermal considerations, including
steady-state temperature drop, and the effect on steady-state hydraulics is
discussed in Chapter 7, entitled Thermal Modeling.
For a more rigorous pipeline sizing, the simulations could be done using transient
simulators. Transient simulators allow changes in parameters such as inlet
flowrate and outlet pressure as a function of time, and calculate values for the
outlet flowrates, temperatures, liquid holdup, etc. as a function of time. If the line
is operating in slug flow, the line size calculated from the transient model might
be different from that calculated from a steady state simulator.
These
considerations will be discussed later in Chapter 8, entitled, Transient
Modeling.
This chapter is arranged as follows. Sections 6.1 and 6.2 discuss some general
guidelines for pressure drop and fluid velocity design criteria (The third design
criteria, flow regimes are discussed in Chapter 5). Section 6.3 presents some
general guidelines to follow when using steady-state simulators.
6.2
For plant piping, rule of thumb values for pressure gradients, such as a
frictional gradient of 0.2-0.5 psi per 100 ft. of equivalent length, are generally
used.
H-0806.35
6-1
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The pressure drop in the line should be compared with the allowable pressure
drop. It should be pointed out that pressure drop is not always a maximum at
the highest flowrate. If the pipeline contains inclined or vertical elements, it is
possible that the highest pressure drop may occur at a low flow condition due
to high elevational losses at low flows.
6.3
Velocity
The velocity in multiphase flow pipelines should be kept within certain limits to
ensure proper operation. Operating problems can occur if the velocity is either
too high or too low, as described in the following sections. It is difficult to
accurately define the point at which velocities are "too high" or "too low. This
section of the guide will try to quantify limits, but these limits should be
considered as guidelines and not absolute values
For the maximum design velocity in a pipeline, API RP-14E recommends the
following formula:
Vmax =
H-0806.35
C
Pns
6-2
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
( Pg V sg ) + ( PlV sl )
Vm
6-3
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Low velocities may cause terrain induced slugging in hilly terrain pipelines
and pipeline-riser systems.
It is not possible to give a simple formula quantifying the velocity when the
phenomena discussed above will occur. The minimum velocity depends on many
variables, including: topography; pipeline diameter; gas- liquid ratio; and
operating conditions of the line. A ballpark value for the minimum velocity
would be a mixture velocity of 5-8 ft/sec. The actual value of the minimum
velocity can only be quantified by simulation of the system using the methods
discussed.
6.4
H-0806.35
6-4
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
6-5
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Although it is possible to get good estimates of the phase equilibrium for 3-phase
(gas-oil- water) systems, the available software does not allow rigorous simulation
of three-phase flow. The models present in steady-state simulators can only do
two-phase (gas- liquid) flow calculations. Steady-state simulators average the
properties of the liquid hydrocarbon and liquid water, and use that average in the
two-phase flow methods. Volumetric averaging, however, may not give good
values for the viscosity and surface tension of the mixture. If the oil and water
form an emulsion, the viscosity estimate may be off considerably using simple
volumetric averaging, because the viscosity of an emulsion can be as much as 50
times as high as the viscosity of the oil or water. If it is likely that an emulsion
will form, the Woeflin method, which is available in steady-state simulators,
should be used to estimate the viscosity of the emulsion.
For lower gas-oil ratios, the choice of models is more difficult. Compositional
models should give more accurate phase equilibrium results, but the physical
property estimates from the compositional models may not be as good as the
black oil model. (Section 4 illustrates this point.) As a result, it cannot be stated
categorically that either the black oil model or the compositional model is
superior for low gas-oil ratio systems. General practice with Steady-State
Simulators has been to use the black oil model for lower gas-oil ratio streams.
If tests of the phase equilibrium and physical properties have been done as part of
the wellstream analysis, steady-state simulators allow the users of the black oil
model to adjust the model predictions for solution GOR, densities, and liquid
viscosity to match experimental values. The pipeline predictions after PVT
matching should be considerably better than those obtained with use of the
standard correlations.
6.4.2
H-0806.35
6-6
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
If the velocities in the line are high, the uphill and downhill holdups may be close.
As the mixture velocity decreases, there will be an increasing difference between
uphill and downhill holdups.
The following table illustrates how sensitive the liquid holdup is to mixture
velocity at various angles of inclination from horizontal. The feed stream is a
gas-condensate with about 4 bbl/MMSCF of liquid present. (The values shown
are predictions of the OLGAS model.)
Table 6-1:
PIPELINE
INCLINATION
ANGLE,
DEGREES
2.7
4.1
5.4
8.1
16.2
-2.0
0.0041
0.0053
0.0064
0.0091
0.0115
-1.0
0.0052
0.0068
0.0085
0.0108
0.0122
-0.5
0.0068
0.0087
0.0108
0.0124
0.0126
0.0
0.0224
0.0218
0.0198
0.0156
0.0131
0.2
0.5797
0.4134
0.2249
0.0179
0.0134
0.5
0.5961
0.4988
0.3846
0.0317
0.0135
1.0
0.5997
0.5000
0.4314
0.3023
0.0144
2.0
0.6009
0.5024
0.4337
0.3428
0.0158
Using the values in the table below, a comparison of two models for a given
section of a pipeline has been made. In the first model, the pipeline segment
consists of two equal length sections of -0.5 degree and +0.5 degree each. The
second model consists of a single horizontal pipeline segment. The liquid
holdups for the two models are:
H-0806.35
6-7
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Table 6-2:
MIXTURE
VELOCITY, FT/SEC
HOLDUP FOR
HORIZONTAL
MODEL
2.7
4.1
0.3015
0.2538
0.0224
0.0218
5.4
0.1977
0.0198
8.1
0.0221
0.0156
16.2
0.0131
0.0131
The liquid holdups are far apart at low velocities and are the same at higher
velocities. This comparison makes two points:
1. The pipeline profile must be realistic if the calculations of liquid holdup and
pressure drop are to be accurate.
2. Low velocities cause severe problems in prediction of the pipeline
performance.
For very low velocities, it would be necessary to know the pipeline elevation
profile within an accuracy of about one pipe diameter in order to get accurate
holdup predictions. This is generally not practical.
In many cases, the pipeline topography is not known when the preliminary
pipeline sizing calculations are run. Frequently, in offshore pipeline design, the
designer only knows water depths at subsea wells or platforms. Instead of
assuming a straight- line pipeline profile, it is recommended that the designer add
some terrain features to the pipeline profile to simulate hills and valleys that are
inevitably present in the actual profile.
To improve the accuracy of the simulation, many calculation segments should be
used in simulating the pipeline. Increasing the number of calculation segments
always improves the accuracy of the simulation, but it increases the computer
simulation time. The number of segments required depends on how rapidly the
temperature, pressure and holdup are changing in the pipeline. For a system with
rapid changes in pressure, e.g. flare systems; the number of calculation segments
H-0806.35
6-8
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
should be greater. If the temperature and pressure are changing slowly, a coarser
grid can be used to simulate the pipeline.
6.4.3
Interpretation of Results
When a multiphase simulator is run, the interpretation of the results can be
difficult. The following section provides assistance in understanding steady-state
simulators output, and ensuring that the design criteria for the line (velocities,
flow regime, and allowable pressure drop) are met.
The velocity in the pipeline should be kept within a limited range. Calculation of
the velocities from a steady-state simulators output is not straightforward. The
designers of steady-state simulators chose to include the actual gas and liquid
velocities in their output table rather than the superficial gas and liquid velocities,
which are needed in the erosional velocity calculations. The superficial and
actual velocities are related by simple formulas:
Vsg = Ug (1 H l )
and
Vsl = U l H 1
It is important to note that liquid holdup is generally shown to only two decimal
places in output tables. For gas-condensate lines, if the liquid holdup is below 0.5
percent, the printout will show 0.00 for the holdup. If there is a way to increase
the number of significant digits to be reported (based on simulator), this is
suggested.
A more accurate way of calculating the superficial velocities from the steady-state
simulators output tables which doesn't rely on reading the value for the liquid
holdup is:
Hl =
(U g Vm)
(U g U l )
Vsl = U l H l
H-0806.35
6-9
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Vsg = Vm Vsl
To calculate the C value in the API-RP14E equation, the value of the no-slip
mixture density must be known. The no-slip mixture density can be calculated
from the phase densities shown on the output table and the superficial velocities
calculated above:
ns =
( g Vsg ) + (l Vsl )
Vm
It is worthwhile to emphasize the point that the pipeline design should be checked
at off-design points as well as the no minal design point. For most pipelines,
worst-case conditions for liquid holdup and flow regime occur at turndown
conditions.
Some steady-state simulators allow the user to print a flow regime map based on
the selected correlations for horizontal and vertical flow. Some simulators have
some limitations. For instance, in PIPEPHASE, the flow regime map is printed
only for the last "device" in a "link. If the "link" contains several pipes with
different inclinations, the flow regime map for some of these sections may be
quite different from the map at the last "device. The only way to print the flow
regime map at specific points along the line is to make these points ends to
"links".
Most simulators also print the flow regime predictions for each pipeline segment.
The printout shows the predictions of the multiphase flow regimes and in some
cases, the correlation/model used (e.g. Beggs and Brill and Taitel and Dukler,
etc.).
Once the flow regime is determined, the designer needs to decide if this flow
regime is acceptable. This decision is more difficult than it may appear. Here are
some general rules/practice methods to follow:
H-0806.35
Ideally, the flow line should not be in the slug flow regime. In practice, it may
be very difficult to design a line to avoid slug flow under all anticipated flow
conditions. The only variables the designer can change are diameter and
operating pressure; the changes in these variables required to avoid slug flow
6-10
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
6.5
The flow regime analysis may show that the line is in stratified flow. In many
instances, this is an excellent flow regime in which to operate. At low
flowrates, however, slugging may occur in lines predicted to be in stratified
flow, induced by the terrain. Terrain induced slugs are generally much longer
than the slugs in normal slug flow and can cause severe operating problems.
Terrain slugging is discussed in more detail in Section 5.2.2.
If the pressure drop and velocities for lines in dispersed bubble or annular
flow are within acceptable limits, these flow regimes are usually good regimes
in which to operate.
6.5.1
H-0806.35
6-11
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
PIPESIM (Baker-Jardine)
OLGA (Scanpower)
HYSIM/HYSIS (Hyprotech)
H-0806.35
6-12
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
In most cases, different flow models can be used to simulate each individual
flowline within a network. Commercial codes such as Pipesim and Pipeflow also
have add-in features allowing optimization of flow through gathering systems and
optimization of lifetime production from a gas field (or group of fields).
6.5.3
H-0806.35
6-13
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The resulting range of pressure drops will be associated with a range of liquid and
gas flow splits and will cover the range of possible pressure drops in the looped
line.
A further effect which must be checked for in looped systems with different
topographies is the possibility of one branch filling with liquid and acting as a
manometer leg (the liquid head balancing the pressure drop in the flowing line.
As well as increasing the pipeline pressure drop this phenomenon could result in
unexpected liquid slugs being received at the terminal when a subsequent flow
increase expels this liquid.
6.5.4
H-0806.35
6-14
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Neptane
Initial flow
Final flow
Mercury
30 Mmscfd
145 Mmscfd
Initial flow
Final flow
Neptane Node 1
Initial flow
30 Mmfscfd
Final flow
145 Mmscfd
Initial HU
1102 bbls
Final HU
273 bbks
Mercury Node 2
Initial flow
25 Mmfscfd
Final flow
25 Mmscfd
Initial HU
236 bbls
Final HU
236 bbks
Niode 1- Node 2
Initial flow
100 Mmfscfd
Final flow
215 Mmscfd
Initial HU
186 bbls
Final HU
48 bbks
Niode 2 Easington
Initial flow
125 Mmfscfd
Final flow
240 Mmscfd
Initial HU
1198 bbls
Final HU
431 bbks
Easington
Minerva
Initial flow
Final flow
25 Mmscfd
25 Mmscfd
Initial flow
Final flow
Liquid
70 Mmscfd
70 Mmscfd
125 Mmscfd
240 Mmscfd
1734 bbls
6.6
How Steady -State Hydraulic Simulation Fits into the Design Process
Steady-state hydraulic simulation fits under the umbrella of Flow System
Thermal Hydraulic Design and Fluid Behavior. The hydraulic design is also tied
in closely with the thermal analysis, which will be the subject of Chapter 7.
6.7
Information Sources
1) Eaton, B.A., Andrews, D.E., Knowles, C.R., Silberberg, I.H. & Brown, K.E.
"The Prediction of Flow Patterns, Liquid Hold-up and Pressure Losses
Occurring during Continuous Two-Phase Flow
2) Gregory, G.A. & Fogarasi, M. "A Critical Evaluation of Multiphase
Gas-Liquid Pipeline Calculation Methods" BHRA, 2nd Int. Conference on
Multiphase Flow, London, 19-21st June 1985.
H-0806.35
6-15
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
6-16
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
7.0
THERMAL MODELING
7.1
Ambient temperatures
With this information, appropriate heat transfer coefficients, which are then used to
determine the temperature profile in the pipeline, can be calculated. Usually, these
calculations are performed by the multiphase simulation software being use, though they
can be calculated manually.
H-0806.35
7-1
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
In most multiphase simulators, heat transfer rates are calculated using homogeneous
mixture properties. Hence no attempt is made to account for the position of the liquid
and gas in the pipe. This approach has been found to be satisfactory for multiphase
pipelines operating under normal transportation conditions.
In some circumstances the homogeneous heat transfer model may not be adequate. For
instance, when investigating heat flow from a pipe or valves during a rupture or shut- in,
knowledge of the liquid content (and resultant heat transfer characteristics) at different
locations in the pipe may be important. Under some such circumstances, fluid
boundaries can be located by experience or calculation, appropriate heat transfer
coefficients can be estimated, and heat loss rates calculated by hand methods. For
others, it would be necessary to use a sophis ticated transient model.
7.2
The relative vapor/liquid split in the separator or flowline (which may affect the sizing
of the flowline or the separator).
The temperature profile at the inlet to the flowline affects the type of insulation and burial
requirements. If the temperature is high, then burial may be required to prevent upheaval
buckling. Conversely, if WHFT is very high, it may be necessary to cool the flow stream
before it enters the flowline.
This design guide discusses the factors, which affect the wellhead temperature, such as
watercut, flow rate, and completion details.
H-0806.35
7-2
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
In the following, heat transfer within the tubing and the casings is not rigorously
calculated. Instead, the casings and annuli are approximated to an equivalent thickness of
cement (or another insulation material). This method provides accurate results
considering that information on the heat transfer through the completion details and the
thermal properties of the surrounding rock may not be readily available and/or of
questionable accuracy.
As an alternate, an unsteady-state well simulation program could be used to rigorously
model the temperature loss through the tubing, casing and the annuli with time. Such a
program can calculate the temperature profile in the well during start-up and as a result of
flow rate changes. Some believe that such a program should not be used to calculate
steady-state wellhead flowing temperatures but should only be used to give an indication
of the warm up rates or the thermal response to changes in the operating conditions such
as flow rate.
This section describes the method used to calculate the WHFT. The description of the
calculation will allow the reader to determine which factors most influence the WHFT.
Calculated Well Heat Transfer
The heat lost from a flowing well into the surrounding rock matrix is by time-dependent
heat transfer. As in the case of a flowline, the overall heat transfer coefficient, which
determines heat loss from the produced fluids, is obtained from a series combination of
three components:
The heat transfer coefficient for radiation is not included as its influence is usually small
compared with the other coefficients.
Heat Transfer Through Completion
The calculation of the heat transfer rate through the layers becomes quite complicated
when a large number of casings are employed. The numerous casings, annuli and cement
thicknesses each contribute to the overall resistance to heat transfer. To simplify the
H-0806.35
7-3
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
2T
[ln( D j +1 / Dj ) / K j ]
For a single layer of cement with an outer diameter of Dc, the equation becomes:
Q=
2K c T
ln( Dc / Dci )
Since Dc is the only unknown and the numerator is the same for both equations the
equations can be rewritten as below:
Dc = Dco exp( K c (ln DJ + 1 / D j ) K j )
The heat transfer coefficient through the completion can now be expressed (based on
tubing inside diameter) as:
hcoml =
2 Kc
Di ln( Dc / DCI )
At the bottom of the well there is normally only one tubing, one casing and one thickness
of cement (and one annulus). However, nearer the top of the well there may be numerous
casings, annulus, different annular fills and cement thicknesses.
Internal Heat Transfer
The internal heat transfer coefficient, hi, assumes the existence of homogeneous twophase flow. This assumption is usually acceptable since the resistance of the internal film
to heat transfer is low compared with the resistance of the casings and the external
medium.
The expression used for estimating the film coefficient is the same as that recommended
for single-phase turbulent flow.
H-0806.35
7-4
1-Dec-00
hi =
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
0.023k m
x Re 0.8 x Pr 0.3
Di
The physical properties are derived on a phase volume weighted basis. The subscript m
denotes a mixture property. For example, the mixture viscosity is determined by the
following:
m = ll + g (1 g )
(Since heat is usually lost from well fluids the exponent for the Prandtl number is set to
0.3.)
External Medium
Heat transfer to the earth is essentially transient radial heat conduction, with the effective
heat transfer coefficient varying with time. While the analysis of the problem is
relatively straightforward, a simple explicit solution is not available. However,
approximate asymptotic solutions for short and long flow times can be derived.
The effective heat transfer coefficient, referred to the tubing inner diameter, is given by:
hearth =
2 ke
Dif (t )
Steady state simulation software able to perform the above calculations may not provide
an accurate prediction of WHFT during start-up. Transient wellbore simulation software
should be used to calculate a more accurate well warm up time.
For flowing times of the order of 10,000 hours, the system can reasonably be assumed to
be approaching steady state. For times between 500 and 10,000 hours the well will
probably be approaching steady state, and the steady-state value will be a reasonable
approximation.
H-0806.35
7-5
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Geothermal Gradient
The undisturbed temperature of the surrounding soil will vary from the reservoir to the
sea/surface. The soil temperature is important since it affects the rate of heat loss from
the fluids. However, the geothermal gradient is not usually known and it is customary to
assume that the (unaffected) temperature varies linearly between the reservoir and the
seabed.
Overall Heat Transfer
Using the coefficients calculated above the overall heat transfer coefficient can now be
determined:
1 1
1
1
= +
+
u hi hcomp hearth
The heat transfer coefficients are based on the inside area. The heat loss from the fluid is
therefore:
Q = UA T
U is the overall heat transfer coefficient (OHTC). Both U and the temperature difference
vary up the tubing and therefore a calculation is best solved by computer simulation.
7.3
H-0806.35
7-6
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
(Tin Tout ) 1/ M
Since Q and Cp can be taken as constant (or relatively insensitive to the flow rate),
increasing M (the mass flow rate) will reduce the temperature loss in the tubing.
Therefore, increasing the mass flow rate will resulting in an increase in the WHFT.
The equation above gives only a rough indication of the effect of flow rate since it fails to
account for the effect of the geothermal gradient. At very low flow rates the wellhead
temperatures tends towards the sea temperature as the fluid temperature tends towards the
geothermal gradient.
Effect of Water Cut
Increasing watercut increases the specific heat and mixture density of the fluid resulting
in an increase in the WHFT. With reference to the equation above, by assuming that the
heat lost from the well (Q) and the mass flow rate (M) to be relatively constant,
increasing water cut will raise the specific heat (Cp) which will reduce the temperature
loss and therefore result in warmer WHFT.
Effect of Gas Oil Ratio (GOR)
The gas phase in an oil well typically provides a relatively small part of the overall mass
of the mixture. Therefore a higher GOR only makes a minor contribution to the overall
enthalpy of the system. Since the rate of heat loss is dominated by the casings and the
external medium, the increased velocity in the tubing has little effect and the heat loss
remains reasonably constant. The effect of the gas oil ratio is less significant than the
watercut or the flow rate.
For example, predictions of WHFT for a range of GORs from 100 to 2,000 SCF/STB
have shown that, for a given flow rate, the WHFT only increased by roughly 2F (1C)
with increasing GOR. The range in WHFTs reduced with increasing flow rate and at 10
KBLPD, the predicted WHFT only varied by 0.13F between 100 and 2,000 SCF/STB.
Effect of Lift Gas
Lift gas greatly complicates the heat transfer model. To rigorously model heat transfer
would require counter flow calculations for the lift section.
For long well flowing times the overall heat transfer is dominated by the heat transfer
through the earth. As a result, the WHFT will be reasonably accurate without
H-0806.35
7-7
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
considering countercurrent flow, provided the mixture temperature of the gas lift gas and
produced fluids is calculated. In this simplified approach the temperature profile takes
place at the gas lift injection point.
Reliance in Predicted Results
As mentioned above, the calculation of the WHFT has a major impact in the design of the
topsides equipment such as the separator and the heat exchangers.
BP has performed a number of brief studies comparing actual WHFTs against predicted
WHFTs for gas wells (West Sole & Bruce) and oil wells (Gyda & Ula). In general, the
predicted results differed from the actual results by less than 5F. In most cases, these
calculations were performed using reasonably accurate measurements of flow rate, water
cut, and bottom hole flowing pressure and temperature taken while testing the well.
In most cases, the WHFT will be calculated using sketchy data from the appraisal wells,
which may not be suitable for the whole development. Usually, the most unreliable piece
of data is the reservoir temperature and the Bottom Hole Flowing Temperature (BHFT);
or FBHT (Flowing Bottomhole Temperature). This not only affects the initial enthalpy
of the wells fluids but also affects the geothermal gradient, which is usually assumed to
be linear between the reservoir and the surface.
Care must be taken when predicting WHFTs for gas wells since, although software
usually accounts for expansion cooling in the tubing, it cannot account for the cooling
between the reservoir and the perforations of the tubing. This may lead to the prediction
of a high WHFT.
The thermal conductivity of the surrounding rock also has a significant influence on the
calculated WHFT. Usually only limited data on the rock type is available and the thermal
conductivity is often assumed to be constant. It is advised that when calculating WHFTs,
the sensitivity to reservoir temperature, flowrate, water cut, rock thermal conductivity are
determined to provide a range of temperatures.
Unsteady-State Predictions
Before start-up, the rock surrounding the completion will be at (or close to) the
geothermal gradient. As the well starts to produce, the surrounding rock will start to
warm. Initially, there is a high temperature differential between the fluid and the
H-0806.35
7-8
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
surroundings and heat loss is high. With time, heat loss will reduce as the rock warms
and the WHFT will increase. The rate of this increase is dependent on the flow rate.
Well start-up can be of particular importance. Before the well is started- up, the
temperature of the fluids is at the geothermal gradient. On start- up, cold fluid will be
produced; if the fluids pass through a choke at the wellhead then the temperature
downstream of the choke may be very low since the initial temperature upstream was
low. This may result in hydrate formation with the potential for forming a blockage.
It is therefore important to know how quickly the well would warm to determine the
length of time that methanol/hydrate inhibitor injection will be required.
When production commences from a well the WHFT will rise quite quickly. Usually, the
wellhead temperature will approach the steady-state value within two residence times of
the well. Typically this is no more than a few hours.
7.4
Pipelines
The temperature of the gas and liquid phases can be computed by carrying out an
incremental heat (energy) balance along the pipe. This balance can be calculated by hand
or (probably more accurately and certainly more conveniently) by using a commercial
software package, which contains the proper physics simulations.
7.4.1
H-0806.35
7-9
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
A large number of temperature profiles could be reproduced here for the range of
conditions likely, but all are of the same type and it should be possible to express them by
a relatively simple mathematical relationship. The simplest possible expression of the
heat loss equation is:
Q = UA T
and in a discretized form, representing the rate of heat loss with distance:
dQ UD(t t c) dL
where Q is the heat flux, U the overall heat transfer coefficient, D the pipe diameter, T
the fluid temperature, A the pipe surface area, T0 the ambient (sea) temperature and L the
distance along pipeline form the inlet.
The equivalent change to the gas enthalpy is given by:
dQ = MCpT (due to reduction in enthalpy)
where m is the mass flow of fluid and Cp the fluid heat capacity. Combining the above
equations gives an expression for rate of change of temperature with distance:
dT / dL = UD( t To ) / mCp
Equation 4.6.4 can be integrated to give temperature as a function of distance:
ln[(T I 0 )/( 0 )] = UDL/mCp
where TI is the fluid inlet temperature. This can be simplified by collecting a number of
terms together for both fluids and expressing the mass flow in terms of volumetric flow,
area and density leads to:
ln[(T1 T0 )/(T - T0 )] = KL/VD
where K is a correlating constant and V the gas velocity.
For this approach to be valid, plotting the left-hand-side of the equation above against
L/(VD) should result in a straight line with gradient K. Test data has been analyzed in
this way and it has been found that K is, to within acceptable limits, independent of
pipeline length, diameter and velocity. It is only a function of CGR, and a simple
relationship between K and CGR has been derived.
H-0806.35
7-10
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
7-11
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
applied to the gas and liquid phases individually and the results combined to give an
overall value for hi.
For low flowrates a convection model is used for the gas heat transfer coefficient:
hi = 0.13k(r2 g|Tw-T|/m2 T)0.333Pr0.333
Tw is the wall temperature and T the fluid temperature. For low liquid velocities the
laminar heat transfer coefficient is used:
hi = 4k/Dh
7.4.2
Rules of Thumb
At early stages of a project, there may not be enough information to enable rigorous
calculation of the heat transfer coefficient. The following rule of thumb values for heat
transfer coefficients for subsea flowlines can be useful in these instances.
TABLE 7-1:
U Value, BTU/hr/ft2 /degF
Applications
H-0806.35
Wells
Risers
20-40
Buried Pipelines
1-3
3-5
5-50
7-12
1-Dec-00
7.4.3
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Maintain fluids at a sufficiently high temperature to avoid, or limit requirement for, feed
heating at the separation plant.
Marine pipelines and risers are provided with some form of coating for corrosion
protection and these offer varying degrees of thermal insulation. Risers on offshore
platforms are normally coated with coal tar enamel or a very thin layer of fusion bonded
epoxy (FBE). These offer little thermal insulation. Often risers are coated with ca. 12
mm of Neoprene to provide corrosion protection in the splash zone. Where thermal
insulation is required this coating may be used over the full length of the riser. Marine
growth on risers can add substantial thermal resistance.
Pipelines are normally protected against corrosion by FBE or coal tar enamel. For an
offshore pipeline requiring concrete weight coating (to provide negative buoyancy) coal
tar enamel is normally applied to the pipe to provide corrosion protection.
Offshore pipelines smaller than 16 inches in diameter are often trenched to avoid damage
by fishing trawl boards. Even if a pipeline trench is not mechanically backfilled (i.e. not
buried) some natural in- fill of the trench normally occurs so that the pipe becomes
partially buried. Under some conditions a trenched pipeline has to be covered by rocks to
provide stability or to protect it from upheaval buckling. This rock dump ing buries the
line; providing enhanced thermal insulation. On some occasions pipelines are buried
solely to provide enhanced thermal insulation.
In addition to the coatings referred to above, low thermal conductivity insulation
materials can be applied to minimize heat loss from a pipeline. Insulation systems can be
divided into two main categories:
Low strength polymeric based materials requiring protection from the environment by
encasing in an outer sleeve. This is the now- familiar Pipe- in-Pipe (PIP) configuration
where a layer of polyurethane foam (PUF) is sandwiched between the inner carrier pipe
H-0806.35
7-13
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
and a protective jacket pipe. For deep and cold water, this type of construction provides
effective and stable pipeline insulation.
Some pipeline coaters are now offering a system for subsea use consisting of PVC or
PUF foam sandwiched between an ethylene-propylene-diene monomer (EPDM)
corrosion coating and an EPDM outer layer providing protection from the environment.
Neoprene.
EPDM.
BTU/hr-ft-F
W/m-K
0.02
0.035
PPF (solid)
0.13
0.225
PPF (foam)
0.10
0.131
Neoprene
0.15
0.260
Epoxy
0.15
0.260
Notes:
(1) Values depend on water depth, temperature and age.
(2) Aging tests on Neoprene carried out in 1985 showed that the thermal conductivity rises
approximately 50 percent as water absorption occurs.
7.5
7.5.1
H-0806.35
7-14
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
TABLE 7-3:
Maximum
Temperature (C)
Coating
FBE (fusion bonded epoxy)
90
70
50
ca. 100
The coating specialists should be consulted if there is any possibility of operating near the
maximum allowable coating temperature. They can also advise on alternative coatings
for any particular application.
CO2 corrosion of pipelines increases with temperature but reaches a maximum at about
60C. In some cases it is necessary to cool the fluids prior to entry to the pipeline to
reduce corrosion rates to a level consistent with the use of carbon steel pipework.
An alternative to providing cooling offshore is to use corrosion resistant pipe for the
initial sections of the pipeline. At some distance downstream of the start of the line
natural cooling in the sea will reduce fluid temperatures to a level where carbon steel pipe
can be used.
7.5.2
Wax deposition
Hydrate formation
High viscosities, in some crudes, and under some circumstances, in oil/water emulsions.
In subsea and onshore multiphase pipelines, fluid temperatures may drop below the wax
appearance point with the result that small particles of wax form. Some of these wax
particles are carried along in the liquid and some are deposited on the pipe walls. In fact
because the walls of the pipeline are generally colder than the bulk fluid, wax deposition
H-0806.35
7-15
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
onto the pipe wall may occur while the bulk pipeline fluids temperature is still somewhat
above the wax appearance point.
7.6
7.6.1
Pipephase
Several concerns arise when using SimScis Pipephase for heat transfer calculations:
The default velocity of water flowing past a pipeline is 10 miles per hour in Pipephase.
This velocity is generally too high. More typical values are 1 to 3 ft/sec (0.7-2 mph).
The Pipephase viscosity routine does not estimate viscosities at temperatures below 60
degrees F. At lower temperatures, it uses the viscosity at 60 degrees F. This can lead to
errors for pipelines in deep water or cold climates.
The thermal conductivity for saturated concrete is much higher than that for dry
concrete. The saturated concrete value should be used for subsea pipelines with
concrete coating.
Unless a value is entered for Hrad, radiation is ignored in the heat transfer calculations.
For subsea or buried pipelines, radiation is negligible, but it can be a significant effect
for surface flowlines.
The convective heat transfer routines in Pipephase are not very rigorous. Errors in heat
transfer calculations can occur for systems in which convection is the prime source of
heat transfer.
H-0806.35
7-16
1-Dec-00
8.0
TRANSIENT OPERATIONS
8.1
Introduction
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Transient multiphase flow simulators have only been developed recently. The first
widely used commercial program, OLGA (marketed commercially by Scandpower),
began development in about 1983 and has been commercially available since 1990. One
of OLGA's current competitors, PLAC (now called ProFES and marketed by AEA), was
introduced to the market at about the same time. A transient simulator called Tacite,
which was developed by IFP and marketed by SimSci, is also available.
Steady state simulators assume that all flow rates, pressures, temperatures, etc. are
constant through time. Inherently transient phenomena, such as slug flow, are modeled
by use of their average holdups and pressure drops. Transient programs show the
variations in parameters such as pressure, temperature, and gas and liquid flow rates as a
function of time and can model phenomena such as slug flow. Transient simulators more
closely model the operation of pipelines and with more detail than do steady state
simulators.
Transient simulators solve a set of equations for conservation of mass, momentum and
energy to estimate liquid and gas flow rates, pressures, temperatures and liquid holdups
as a function of time. The programs utilize an iterative procedure that ensures that a set
of boundary conditions (such as inlet flow rates and outlet pressures as a function of
time) are met while solving the conservation equations.
Generally, the use of transient simulators will be more appropriate than steady state
programs whenever transient performance is highly important to the design process.
Such situations include startup and shutdown liquid and gas delivery, heatup and
cooldown temperature fluctuation, unsteady flow into receiving facilities (separators and
slug catchers), and pigging. In addition, the current version of OLGA2000 includes the
ability to estimate the flow of all three phases (oil, gas, and water). Particularly at low
production rates, liquid holdups (and the distribution of oil and water in the holdup
liquid) should be more accurately predicted when using three-phase simulation.
Use of Transient Simulators
Because of their utility, transient simulators have been used for a variety of purposes.
These uses include:
H-0806.35
8-1
1-Dec-00
8.2
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Pigging simulation
Operator training
H-0806.35
8-2
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
known about them. It is expected that competition in the market place will increase when
some of the other codes are commercialized.
8.2.1
OLGA
OLGA is a dynamic, one dimensional modified two fluid model for transient two-phase
hydrocarbon flow in pipelines and networks, in which processing equipment can be
included. The code has been developed by the SINTEF/IFE two-phase flow project
which commenced in 1984 and was based on the computer program OLGA 83. This was
originally developed for Statoil by IFE in 1983.
The two-phase flow project was aimed at improving OLGA by expanding an
experimental database from a high pressure 8- inch large scale test facility run by SINTEF
at Tiller in Norway. Extensive testing was carried out by IFE and by the projects
member oil companies which included Conoco Norway, Esso Norge, Mobil Exploration
Norway, Norsk Hydro, Petro-Canada, Saga Petroleum, Statoil, and Texaco Exploration
Norway.
The basic models in OLGA contain three separate mass conservation equations for the
gas, the continuous liquid and the liquid droplets, these are coupled with the interphase
mass transfer terms. Momentum conservation is applied to the gas-droplet field and to
the continuous liquid, hence giving two equations. The mixture energy equation is
written in conservation form, accounting for total energy balance in the system. OLGA
predicts as a function of time the pressure, temperature, mass flow of gas and liquid, the
holdup and the flow pattern. Closure laws are required for the friction factor and the
wetted perimeters of the phases and these are flow regime dependent. The droplet field
also requires an entrainment and deposition model. The two basic flow regimes adopted
are distributed and separated flow. The former contains bubble and slug flow and the
latter stratified and annular flow. The transition between the two regime classes is
determined according to a minimum slip concept.
Due to the numerical solution scheme, the original versions of OLGA are particularly
well suited to simulate slow mass flow transients. The implicit time integration applied
allows for long time steps which is important for the simulation of very long transport
lines, where typical simulation times are in the range of hours to days.
The necessary fluid properties (gas/liquid mass fraction, densities, viscosities, enthalpies
etc.) are assumed to be functions of temperature and pressure only, and have to be
supplied by the user as tables in a specific input file. Thus, the total composition of the
H-0806.35
8-3
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
two-phase mixture is assumed to be constant both in time and along the pipeline for a
given branch. The user may specify a different fluid property table for each branch, but
has to ensure realistic fluid compositions if several branches merge into one.
In 1989 Scandpower A/S acquired marketing rights to the OLGA program and
commercial use of the code increased considerably. The early versions of OLGA had a
simple data input and output file format which made the setting- up of problems very time
consuming. Later versions have user friendly interfaces which greatly enhances the use
of the code, PreOLGA is used for input data generation and PostOLGA is used for post
processing. Fluid properties are generated using a package called PVTOL. The newest
version of OLGA (OLGA2000) has further improved the user interface and has improved
calculation accuracy for high GVF flow situations.
In 1993 Statoil acquired the Tiller facility for 5 years and embarked on a substantial
research and development effort to improve OLGA, areas under development include:
Compositional tracking.
These areas also indicate where there are deficiencies in the present code.
In addition, OLGA has been interfaced with the Fantoft D-SPICE multi-compositional
dynamic process plant simulator to improve the modeling of equipment; however, this
interface only provides a dynamic exchange of data between the codes and is hence not a
fully integrated model.
Some of the other original two-phase flow project sponsors have also developed OLGA
for their own use, notably Conoco who have developed improved thermal modeling and
pigging simulation in their version called CONOLGA.
8.2.2
PLAC
The PipeLine Analysis Code (PLAC) was originally developed from the nuclear reactor
safety code TRAC (1986) under a four-year program sponsored by BP Exploration and
the UK Offshore Supplies Office. The major parts of that work were to remove the
redundant reactor specific components and to convert from steam/water physical
properties to multi-component hydrocarbon properties. More suitable models for
H-0806.35
8-4
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
interfacial friction and flow regimes were also introduced. The first release of PLAC was
available for use by AEA Petroleum Services and BP Engineering in mid 1990 though
commercialization of the code was not realized until the autumn of 1992. BP and AEA
conducted numerous validation tests on the code with good and bad experiences. Like
OLGA, the original versions of the code were cumbersome to use but ease-of-use has
been greatly improved by the development of interactive pre and post processors. PLAC
is now marketed by AEA.
PLAC solves mass, momentum, and energy equations for each phase using a onedimensional finite difference scheme. Six equations are solved: gas and liquid mass and
momentum conservation, total energy conservation, and gas energy conservation. Unlike
OLGA, PLAC does not have a separate equation incorporating a droplet field in the gas
stream. Appropriate flow pattern maps and constitutive relationships are provided for
wall and interfacial friction and heat transfer, and a model for multi- component phase
change is included. PLAC has flow regime maps for vertical and horizontal pipes and
switches to vertical flow if the angle of inclination is above 10 degrees. The horizontal
flow pattern map is based on the method of Taitel and Dukler (1976).
The fluid physical properties are calculated from a user-supplied mixture composition
using an internal PVT package, however this generates a table of properties similar to
OLGA, and hence still relies on simple equilibrium phase behavior predictions. PLAC
can only use one composition in a network. The current version of the code has the
capability to handle pipes, tees, and valves and is able to predict a range of phenomena
including flow rate and/or thermal transients, severe slugging, shutdown/restart problems,
and pipeline depressurization.
8.2.3
WELLTEMP
WELLTEMP was designed to examine temperature transients caused by flow in wells. It
can also be used to examine temperatures in pipelines. It handles heat flow through
tubing, casings, annuli, cement and formation. Various operations can be modeled such
as injection, production, forward and reverse circulation, shut-in and cementing. It
assumes steady state flow rates; although different flow periods can be entered (steady
within each period). It can handle dual or even triple completions (and co-bundled
flowlines). In dual completions it would allow injection down one string and production
back up the other.
It can be used to examine surface and downhole tubing temperatures during drilling or
simulated treatments. This is useful when considering fluid properties, cement
H-0806.35
8-5
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
characteristics etc. It can also help when considering corrosion, wax deposition, or
hydrate formation. It can be used to examine cool down times in pipelines. It could also
be used to refine the temperature profile in the annulus and wellbore during gas lift
operations.
WELLTEMP, marketed by Enertech Engineering and Research Company of Houston, is
primarily useful for evaluating warm- up of wells and flowlines. WELLTEMP can
usually run much faster than PLAC or OLGA and has been used to investigate
sensitivities before performing a detailed PLAC or OLGA simulation. The simple
composition treatment in WELLTEMP is a limitation to detailed transient two-phase
8.2.4
Tacite
(Information on Tacite to be added)
8.3
Startup
In normal operation hydrates can be controlled by maintaining the flowing temperature
above the hydrate formation point and/or by injecting a sufficient volume of inhibitor.
The required dosage is determined by using a appropriate program or the GPSA charts
for the prevailing conditions. During shut-down and start- up, conditions can occur which
increase the potential for hydrate formation, and therefore require additional quantities of
inhibitor. Figure 8.3-1 shows a typical hydrate disassociation curve. A pipeline
pressure/temperature profile is superimposed as line A representing conditions in a
pipeline from a subsea template to the slug catcher on a host platform, where hydrate
inhibition may not be required under normal operating circumstances. In some cases the
pipeline may be long and the fluids may cool to seabed conditions where some inhibition
may be required. This could also occur as a result of lower flow rates, and is shown as
curve B. Hence, in normal operation steady state analysis can be used to determine the
amount of inhibitor required.
H-0806.35
8-6
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-7
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-8
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The calculation has three main elements. These are: depressurization rate, JouleThompson cooling, and heat transfer from the surroundings.
The depressurizing rate is a critical aspect. The faster the depressurizing the lower the
temperature that will be experienced.
The venting rate can be fixed (e.g. 100 MMSCFD) by using a control valve and a mass
flow rate meter. Alternatively, the system can be vented through an orifice. In most
cases the flow through the orifice is critical, and therefore the depressurizing rate is
independent of the downstream pressure.
As the pressure decreases, the temperature of the system will drop due to JouleThompson cooling (auto-refrigeration). The expansion of the gas is normally considered
to be isentropic. As the gas expands it does work and the energy required to do this work
is removed from the gas in the form of heat. For an isentropic process, the autorefrigeration is not dependent on the duration of the depressurizing, only the physical
properties of the fluid. However, as discussed below, the process is not truly isentropic.
Depressurization of a pipeline system usually requires many hours, and in some cases
days. During this period, the heat input from the surroundings may be (for uninsulated
systems, usually will be) significant.
The Joule-Thompson effect is not time-dependent and is only dependent on the expansion
of the gas in the system. Heat transfer from the surroundings is time dependent. The
longer the depressurization, the greater the influence of the surroundings whilst the autorefrigeration aspect remains (roughly) constant. Therefore, increasing the depressurizing
time will result in less cooling and, therefore, in a higher minimum temperature.
Depressurizing Rate Calculation
It is usual to depressurize a system through an orifice or a valve. The equation for the
mass flow rate through an orifice is given below. The equation for a valve is of a similar
form with slightly different constants (not given here).
W = Cd KAP 1 M W / zT
where
Cd = Coefficient of discharge
H-0806.35
8-9
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
2
+1
K = [( )[(
)^ (
)]]^ 0.5
R +1 + l
where
P2
2 ( / 1)
<(
)
P1
+1
where P2 = Downstream pressure (N/m2 ).
The occurrence of critical flow greatly simplifies the calculation. If the system is
operating under critical flow then the flow rate is only dependent on the upstream
pressure.
It is normal to consider the depressurizing to be complete when the system pressure falls
below 2 bara. For a downstream pressure of 1 bara, critical flow exists down to pressures
of 1.83 bara (g = 1.3) and so therefore critical flow is usually assumed for the whole of
the depressurizing. Critical flow may not exist at the lower flow rates where the
downstream pressure is significantly greater than atmospheric,
While the pressure ratio satisfies the last equation, the flow rate (W) is independent of the
downstream pressure.
H-0806.35
8-10
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Heat Transfer
The temperature of the fluid is determined by the effect of Joule-Thompson cooling due
to isentropic expansion of the gas and warming of the gas by heat transfer from the
surroundings. The conservatism of the equation is described below:
The assumption of isentropic expansion means that the gas loses the maximum amount of
energy (i.e. heat) and therefore results in a conservative prediction of the minimum
temperature.
When the gas temperature drops below the ambient temperature the pipeline is warmed
by the surroundings. The influence of the surroundings must be reasonably estimated so
as to calcula te a conservative (i.e. low) minimum temperature.
The calculation of an accurate internal heat transfer coefficient is important since it
greatly influences the prediction of the minimum temperature. The method used must
calculate a conservative (i.e. low) heat transfer coefficient for at least the period that the
system is at its minimum temperature.
The correlation used to determine the heat transfer coefficient will depend on the type of
system and the method of depressurizing. Details of the equatio ns used to calculate the
internal heat transfer coefficient are given below.
H-0806.35
8-11
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Table 8-1:
Heat
Transfer
Equations
The calculation of an accurate internal heat transfer coefficient is important since it greatly influences
the prediction of the minimum temperature. The method used must calculate a reasonably
conservative (i.e. low) heat transfer coefficient for at least the period that the system is at its
temperature.
The correlation used to determine the heat transfer coefficient will depend on the type of system and the
method of depressuring.
The internal heat transfer coefficient is calculated using one of the following equations:
Stream Line Flow (Re < 2100)
If the flow is laminar (Reynolds number < 2100) then the following equation should be used:
1.86 k m d i u m m
hi =
d i m
c pm m
d i
1/ 3
1/ 3
)]
log 10 Re
[Process
Heat
H-0806.35
8-12
1-Dec-00
Gra =
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
d i3 m2 gT
m
0.023k m
hi =
di
du m m
c pm m
k m
0 .4
m = l + v (1 )
The velocity at which the heat transfer coefficient is to be calculated should be 75% of the exit
velocity. This allows for the fact that there is a velocity profile through the system. If possible
the velocity should also be based on the average velocity for the time step.
8-13
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
temperature. If the pipe is located in the sea, the overall heat transfer coefficient will be
higher and the minimum wall temperature will accordingly be higher (assuming, for this
comparison, that the air and sea temperatures are equal)
Minimum Pipe Wall Temperature
The objective of most depressurizing studies is the calculation of a minimum pipe wall
temperature, or the calculation of a depressurizing rate which prevents the system from
falling below the minimum tube wall temperature.
In a gas-only system the heat transfer rate between the gas and the pipe wall will be low,
and consequently the temperature difference between the fluid and the tube wall will be
relatively high. In order to be conservative, it is usual to assume that the fluid temperature
and the tube wall temperature are identical throughout the depressurizing. This
assumption will always result in the calculation of a conservative (i.e. low) minimum
tube wall temperature since it ignores the existence of a temperature profile between the
fluid and the wall.
In a two-phase system, the heat transfer rate between the liquid and the pipe wall will be
significantly higher than the heat transfer rate between the gas and the pipe wall. It is
reasonable to assume that where pools of liquid occur (in dips etc.) the high heat transfer
rate will result in the pipe wall cooling to the same temperature as the fluid. However the
effect of the fluid being warmed by the pipe wall is ignored and hence a conservatively
low pipe wall temperature is predicted.
Simultaneous Riser and Topsides Depressurizing
In most studies the riser and topsides pipework will be depressured simultaneously.
The majority of the length of the riser will be immersed in the sea and have the benefit of
the sea as a heat source. Riser pipework is generally much simpler than the Topsides
pipework.
The topsides pipework is more complicated (e.g. a large number of branches etc.), and
located in the air. The poor heat transfer properties of air coupled with the lower
minimum ambient temperature mean that the topsides pipework will drop to a lower
temperature relative to the riser. Because of this difference in heat transfer rates the
topsides and the riser should be modeled separately.
H-0806.35
8-14
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-15
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Blowdown Philosophy
A blowdown philosophy document defining relevant cases and the start conditions from
which blowdown occurs should be developed in the early stages of any design.
Code Requirements
API RP 521 provides general guidelines for determining the need for emergency
blowdown. There are, in fact, no mandatory blowdown requirements. However,
emergency blowdown has become accepted in the industry as being necessary for all
offshore installations, other than for some unmanned platforms. The BP Code of Practice
CP 37 clearly defines this philosophy.
CP 37 sets out the three basic reasons for providing emergency blowdown facilities as:
1. To reduce the risk of vessel or pipeline rupture in a fire.
2. To minimize the fuel inventory which could supply a fire.
3. To minimize the uncontrolled release of flammable or toxic gas.
Emergency blowdown facilities are not required for subsea pipelines. The platform
topsides section of the pipeline, including pig launcher/receiver and all pipework on the
platform side of the ESDV should be blown down through the platform blowdown
system. For pipelines fitted with subsea isolation valves (SSIVs) there is also a
requirement to provide for emergency blowdown of the riser and subsea pipework
situated between the SSIV and the top of the riser.
Effects of Blowdown
Blowdown of the contents of a vessel, pipework, or sections of a pipeline from high
pressure results in auto-refrigeration of the contained fluid due to expansion of the gas
and evaporation of the liquid. The resultant fall in the temperature of the pipe or vessel
wall metal can result in the need for costly materials to avoid brittle fracture. In addition
low fluid temperatures can result in the formation of hydrates, ice, or even solid CO2, and
these can lead to blockages.
Blowdown Times and Start Conditions
Blowdown must be at such a rate that equipment subject to high temperature does not
fail. API RP 521 sets out the basic requirements for blowdown times and these are
H-0806.35
8-16
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
8.5
8.5.1
H-0806.35
8-17
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
If it is assumed that no liquid is expelled from the pipeline outlet before steady state is
reached (the case for all but extremely unusual conditions or topographies) this is a
simple calculation. The results are most useful plotted as a Line Fill Nomograph as
shown in Figure 8.5-1, an example for a 125 km long, 24- inch diameter line.
The sloped section of the curve is the line filling, the gradient being proportional to the
gas flow rate. The flat section of the curve occurs once steady state has been reached. In
the case shown, the highest steady state holdup (25,000 BBL) occurs at a gas flow rate of
75 MMSCFD exceeded the existing slugcatcher capacity by nearly 20000 BBL.
However continuous operation at this low gas flow rate would be required for a period of
7 days to reach this liquid content.
Fill-time nomographs may also be of use in the general operation of lines for reckoning
of liquid contents following a series of flow rate changes as filling lines of the
appropriate gradient can be extrapolated for cases where the pipeline is already partially
full of liquid.
H-0806.35
8-18
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
In general, decreases in gas flow rate, however dramatic, will not cause liquid production
(unless accompanied by a significant decrease in outlet pressure). Instead, liquid
production will cease following the reduction in flow until the liquid content of the
pipeline has been reached. Liquid production will then resume at a flow rate
commensurate with the gas flow and CGR. The period required for liquid accumulation
can be quickly estimated from a fill-time nomograph.
Liquid production during and after a flow increase will depend upon the initial liquid
content of the line, the initial and final flow rates, the rate at which the flow rate is
increased and the topography of the line.
8.5.2
8.5.3
Effect of Initial and final gas flow rates on Liquid Delivery Rates
Flow rates are important in determining the size of the peak in liquid production as
described above but will also add to the background liquid production rate during the
transient and the rate at which a peak in liquid production rate occurs. Various simple
calculation and estimation methods have been devised to predict liquid production during
transients based on initial and final flow rates and liquid holdups. Gregory (1994)
suggests that the two extremes in the time for a slug of liquid to appear and the rate at
which it appears can be estimated from the residence times of the liquid and gas phases in
the pipeline:
H-0806.35
8-19
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
f = L (1 - e) / V
f = L e / V
f is the residence time of the excess liquid in the line, V?the gas superficial velocity,
V the liquid superficial velocity, L the length of the pipeline and e the average voidage.
The first expression assumes the excess liquid moves as a slug and travels at the average
gas velocity. The second assumes that the liquid moves as a layer or film traveling at the
average liquid velocity. This approach then estimates the liquid production rate, Q,
during the transient as:
Ql = (vi - ve ) /
where vi is the initial liquid content of the line and ve the final liquid content. To Ql must
be added the steady state liquid production rate during the transient which is estimated as
the liquid flow rate associated with the increased gas flow.
This method was compared with field data for two gas-condensate lines (one 18-inch
diameter and 30 miles long, the second 20- inch diameter and 67 miles long) by Cunliffe
(1978) who obtained good results basing the residence time on the average liquid velocity
and predicting liquid holdups with the correlation of Eaton et al. (1967).
A slightly different approach was taken by Bevan (1992) who used analogy with a
conveyor belt to predict liquid production during and following flow rate changes.
Conveyor speed is dependent upon the inlet gas velocity and this determines the transport
of liquid out of the line. The amount of liquid "on" the conveyor at each point along its
length is determined by the gas velocity when that section of "conveyor belt" entered the
line. Changes in flow rate are thus modeled as fronts of different holdup which travel
through the line at a rate dependent on the instantaneous inlet gas velocity. Bevan backcalculated from the Eaton and other correlations and comparisons with experimental data
to produce holdup correlations for this model.
Both Gregory/Cunliffe and Bevan models have been shown to provide acceptable
estimates of transient liquid production but cannot easily allow for the effects of pipeline
topography or ramped increases in flow rate.
Again, a transient simulator will generally give a more accurate understanding of fluid
behavior during turn-down or ramp-up and, depending on the simulator being used and
on the fluid being modeled, may yield more accurate liquid and gas volumetric estimates.
H-0806.35
8-20
1-Dec-00
8.5.4
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
8.5.5
H-0806.35
8-21
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
represented on this plot by a line with maximum gradient equal to the processing capacity
(note however that this assumes that the liquids processing plant throughput can be
increased instantaneously). Since liquid can only be processed when it is present in the
slug catcher/separator the processing line follows the liquid production curve initially. At
point A, where the transient liquid production exceeds the processing capacity the
processing rate diverges at its maximum gradient. The processing line rejoins the liquid
production line at a later time, after the pipeline has reached steady-state, when the
backlog of liquid has been processed.
H-0806.35
8-22
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The most widely used method in these circumstances is to routinely pig the line using
spheres. This method is most suited to long pipelines with a high steady state liquid
holdup flowing low CGR fluids. The pigging interval is planned so that the liquid
accumulated in the pipeline is a conservative fraction of the available slugcatcher volume
(this interval can be estimated from the line- fill nomograph such as Figure 8.5-1). The
line is only allowed to reach steady state at high flow rates (when the holdup is low) and
any flow ramp can be accommodated as the line never contains more than the slugcatcher
volume of liquid. If high CGR fluids are flowed, sphering as a means of inventory
control may not be practical, as the pigging interval will be too short to allow normal
pipeline operation. Other disadvantages of this approach are the need to install pig
launching and receiving facilities, the need to design the pipeline to accommodate pigs,
the risk of pigs sticking in the line due to wax, hydrate or other debris, the need to
transport pigs back offshore following use and the risk of pig launcher and receiver
failures shutting down the terminal.
An alternative approach is to modify the pipeline operating procedures to reduce the
severity of transients. Williams (1996a & b) used transient simulations to define an
operating envelope, see Figure 8.5-3, for an export pipeline and for a proposed infield
flowline, operation of which is complicated by the flowing of fluids ranging from CGR
20 to 200 BBL/MMSCFD. LPDs were calculated from transients ramping the gas flow
from low values to the design maximum based on the existing slugcatcher, marginally
oversized liquid processing facilities and a ramp time of 24 hrs (based on the gas
processing plant turn- up rates). The envelope identified the operating conditions from
which the ramp to maximum flow rate could be performed. Operation outside the
envelope of flow rates requires more time for flow increases and the use of line fill time
nomographs to assess the accumulation of liquid in the line.
H-0806.35
8-23
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-24
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-25
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Mechanistic Model, less that swept out of the pipeline as a liquid film during the sphere
transit time. To determine the sphered liquid volume the growth of the slug in front of
the sphere must be calculated, using the mass balance equation, throughout the spheres
passage through the line.
The process involved with the passage of a sphere through the flowline is shown in
Figure 8.6-1.
H-0806.35
8-26
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
collecting liquid from the preceding film. Liquid leakage passed a sphere that does not
fully block the pipe can also be included.
Thus, a combination of three equations for:
The mass balance at the sphere (if liquid leakage passed the slug is modeled)
The growth of the slug due to the differential velocity between the slug front and
sphere
lead to the required method for integrating the slug length as the sphere passes through
the flowline.
For the special case of a horizontal flowline, zero pressure drop, no liquid leakage passed
the slug and a sphered slug holdup equal to unity, this equation reduces to:
Vprod = Vtot Vout
where:
Vtot = total liquid hold-up in the line prior to sphering
Vout = liquid volume leaving the line during passage of sphere
This approach assumes that the pig travels at the in situ mixture velocity, which is not
always the case in practice. The motion of the pig may vary as a result of the
compressible nature of the fluid in the pipeline, which allows pressure and velocity
excursions to arise from the variations in the pipeline inclination and the friction between
the pig/pig train and the pipeline. For example, the pig may require additional driving
pressure to push the slug over a hill or to compensate for an increasing frictional
resistance as the slug grows in length. If the system is spongy the slug ve locity may
decrease or even stop as the fluid packs behind the pig to provide the additional driving
pressure. In downward sections, head recovery can cause the pig to accelerate and
transfer excess pressure into friction loss.
If the pig stops, then it is possible to observe a stick-slip phenomena caused by static
and dynamic friction effects, where more pressure is required to get the pig moving than
is required to keep it moving. The additional pressure required for re-start can cause the
pig to accelerate due to the lower dynamic friction. This higher pig velocity can de-pack
the line resulting in less driving pressure and a tendency for the pig to stop again.
H-0806.35
8-27
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
These phenomena can have a large effect on the time required to pig a pipeline, on the
dynamic pressures generated, and on the fluids produced. This may affect the design and
operability of the system. In addition, it is sometimes important to predict how long it
takes to re-establish the liquid holdup behind the pig, and hence to re-distribute corrosion
inhibitors.
Simple pigging transients can be investigated using the OLGA and PLAC transient twophase flow codes, both of which allow for a variable leakage past the pig during the
pigging process. The PLAC model is a recent addition and has yet to be properly
validated, but is used here to demonstrate the transient effects caused by pigging
pipelines.
Transient pigging example-pigging a hilly terrain pipeline
Figure 8.6-2 shows the topography of a hypothetical hilly terrain pipeline which is 200
mm in diameter, 10 km long, and includes a 50 m high hill followed by a 50 m deep dip.
A high GOR composition was used and the simulation run to a steady state with an inlet
mass flow of 2 kg/s and a delivery pressure of 10 bara. A pig was launched into the
pipeline at time 20100 s and it was estimated that the pig transit time would be around
2900 s based on the average mixture velocity in the pipe. Figure 8.6-3 shows the pig
position with time, and indicates that the pig motion was not steady; and displaying three
regions of steady motion and two regions of slower motion, and in some cases reversal.
The actual pig transit time is 5900 s, which is almost twice as long as estimated from the
average mixture velocity. This discrepancy results from the fact that the pig slows down
in the uphill section as the pressure builds-up.
H-0806.35
8-28
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
60
Elevation (m)
40
20
0
0
2000
4000
6000
8000
10000
12000
-20
-40
-60
Distance (m)
H-0806.35
8-29
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 8.6-4 shows the holdup profile throughout the pipeline for various times and indicates that
before the pig is launched the holdup in the upward inclined sections is around 35 percent,
whereas in the horizontal and downward sloping sections it is around 2 percent. The pig slug is
shown just after the pig has passed the top of the first hill and indicates that the holdup is around
95 percent, and that the slug is less than 500 m in length. The holdup profile at 40,000 s
indicates that the liquid content of the pipeline is still building-up 4 hours after the pig has been
received. Figure 8.6-5 shows the pressure at the pipeline inlet and indicates that a pressure rise
of 1.3 bara is required for the pig to ne gotiate the first 50 m hill. The liquid full hydrostatic head
would be equivalent to 2.8 bara, hence the generated pressure is consistent with the observation
that the pig slug half fills the upward inclined section.
Time in sec
H-0806.35
8-30
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
8.7
H-0806.35
8-31
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
8-32
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
= fluid density
If the fluid is stopped instantaneously, the maximum head or so called Jacowski
pressure rise results. In a practical system the maximum head may be generated if the
fluid is brought to rest within the pipeline period. For example, consider a pipeline
discharging through an outlet valve which closes fully. The full Jacowski pressure rise
occurs if the valve closes within the time taken for the pressure wave to travel to the other
end of the pipeline and back. This is the pipeline period and is given by:
= 2L / a
where:
= period
L = length of the pipeline
This method may be used to check the potential for over pressuring the pipeline, but may
not give the maximum pressure if networks are involved due to amplification at the dead
ends.
There are several operating changes that can cause flow changes and hence pressure
transients. The most common are valve opening/closing and pump start/stop.
In the valve closure example the shape of the pressure/time variation will depend on the
flow/time change which caused it and thus on the characteristics of the valve. Typical
flow characteristics of ball, globe, and gate valves are shown in Figure 8.7-1. It is
important to note that the last 10 percent of the valve closure has the most significant
effect.
H-0806.35
8-33
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 8.7-1: Relative flow rate against percentage valve opening for typical gate,
ball, and globe valves
If the pipeline period is greater than the valve movement time then the type of valve
characteristic will not influence the total pressure rise, although the rate of pressure rise
will be important to any control devices and the mechanical load on supports. If the
period is less than the valve movement time then the shape of the characteristic will
influence the peak pressure. Special care should be exercised in specifying valves whose
effective closure time is only a fraction of the total movement, e.g. gate valves.
Pressure falls can be as damaging as pressure rises. Often the negative head changes lead
to theoretical pressures less than the fluid vapor pressure. In such cases a vapor cavity
forms. Large diameter lines are more susceptible to damage by reduced pressure
conditions. More commonly vapor cavities create damage by their collapse. This occurs
if there is a situation, which can re-pressure the line. This could be caused by refilling
from an elevated section of line or tanks or by re-starting a pump to continue operation.
Often this refilling occurs at a high rate because the pump discharge pressure is lower
than normal. When the cavity volume reaches zero the effect is of a truly instantaneous
valve closure. These positive pressures can be higher than with normal valve closure and
have step pressure changes. Therefore, they should be avoided if at all possible.
H-0806.35
8-34
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Pump start/stop has the capability of producing similar positive and negative pressure
waves with the attendant damage potential.
From the above, it becomes apparent that the significant factors in creating high surge
pressures are the wave speed and the rate of change of flow. The wave speed is fixed by
the fluids used, although it is sometimes possible to modify the rate of change of velocity
in a transient. This is particularly true of the effective valve closure time. Even valves
with the most suitable closure characteristics create most of the flow change over the last
quarter of their movement. Thus, two part closures, where the first part is fast and the
last part slow can be beneficial.
Where surge is a problem, it is particularly important to avoid smaller bore area ends.
When the surge pressure reaches a pipe junction it propagates along both lines. The
pressure reduces in ratio to the new area divided by the old area and induces a flow by the
pressure difference. At the instant of the wave arriving at the dead end the forward
velocity is stopped creating another surge effect on top of the pressure wave. This can
lead to a potential doubling of the pressure.
One solution to reduce surge pressures is to introduce surge accumulators into the system.
This is particularly effective with pump- generated surges. These accumulators are
vessels which contain a gas pocket connected as closely as possible upstream of the surge
generation device so that when the surge occurs, the excess fluid flows into the vessel
compressing the gas. By removing excess fluid (which is nearly incompressible) the line
pressure is reduced. The volumes and costs of such suppression equipment can be quite
small when the transients are rapid.
Wave speed in single phase flow
Calculation of the speed of pressure waves in the fluid medium is crucial to determination
of pressure surge. For a single phase system the pressure wave velocity is given by:
= (K/)1/2
where:
K = bulk modulus of elasticity
= density of the fluid
Hence for water at 15 C, K = 2.15 GN/m2 and = 1000 kg/m3 the speed of pressure
waves is 1466 m/s. For an ideal gas the bulk modulus is the ratio of the change in the
H-0806.35
8-35
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
pressure to the fractional change in volume and hence depends on whether the
compression is adiabatic or isothermal. For an adiabatic process PV = (constant) and
hence K = P. Since P/ = RT the expression for adiabatic compression of an ideal gas is:
= (K/)1/2 = (RT)1/2
For air at 15 C, = 1.4, R = 287 J/kg/K; hence the wave speed is 340 m/s. The
equivalent bulk modulus of elasticity is 141.7 KN/m2 and the density is 1.226 kg/ m2 .
The pressure wave velocity is modified when boundaries such as pipe walls, free
surfaces, gas bubbles and solid particles are present. The elasticity of the walls of the
pipe through which the fluid is travelling has the effect of reducing the pressure wave
speed by a factor dependent upon the size, cross-section shape, and pipe material. In
addition the method of pipe restraint may also affect the result. The general equation for
the wave speed in a fluid contained within a thin walled pipe of circular cross-section is:
a = 1 / [ (1/K + D/tE)]1/2
where:
D = internal diameter
t = wall thickness
= pipe restraint factor
For materials with a high elastic modulus such as steel or concrete, or for pipelines with
frequent expansion joints, F can be taken as unity. Hence for water at 15C in a steel
pipeline (E=207 GN/m2) of 200 mm diameter and 15 mm wall thickness, the wave speed
is reduced from 1466 m/s for rigid pipewalls to 1374 m/s.
Wave speed in two-phase flow
The wave speed in two-phase flow can be determined by simply replacing the single
phase bulk modulus and density with the two-phase equivalents, i.e.:
a = (Ktp / tp)1/2 for rigid pipewalls
and:
a = 1 / [ tp (1/Ktp + D tE)] 1/2
where:
H-0806.35
8-36
1-Dec-00
(1- )
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
= +
Ktp
KL
Kg
and:
tp = (1- ) L + g
= in-situ gas fraction
Using air and water as an example it is seen that the wave speed is rapidly reduced by the
introduction of only small quantities of gas. For example, a void fraction of only 1
percent is required to reduce the wave speed from 1446 m/s to 119 m/s. The table below
shows the wave speed for vario us gas fractions, also shown is the relative Jacowski head
rise possible compared with the liquid only case. It is hence seen that in two-phase flow
the wave speed and potential surge is much reduced by the combined high
compressibility and high density of the two-phase mixture.
Table 8-2:
Void
fraction
0.000
0.005
0.01
0.1
0.2
0.4
0.5
0.6
0.8
0.9
0.99
0.995
1.0
Two -phase
bulk modulus
(N/m2)
2.19e+9
2.80e+7
1.41e+7
1.42e+6
7.08e+5
3.54e+5
2.83e+5
2.36e+5
1.77e+5
1.57e+5
1.43e+5
1.42e+5
1.42e+5
Two -phase
density
(kg/m2)
1000.0
995.0
990.0
900.1
800.2
600.2
500.6
400.7
201.0
101.0
11.2
6.2
1.2
Wave
speed
(m/s)
1466.0
168.0
119.0
37.7
29.7
24.3
23.8
24.3
29.7
39.4
113.0
151.3
340.0
% potential head
rise
100.0
11.4
8.0
2.4
1.62
1.00
0.81
0.66
0.41
0.27
0.07
0.06
0.03
H-0806.35
8-37
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Topography Effects
8.8.1
H-0806.35
8-38
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Interactions with the terrain may cause additional effects. For example, the liquid
sloshing in a dip may cause perturbations in the gas and liquid flow rates downstream,
giving rise to flow regimes different to those expected from analysis with steady flow
rates. Flow rate fluctuations caused by slugs dissipated in a downward sloping section
may be sufficient to generate slugs in a following downstream section normally operating
in stratified flow.
By taking account of the hysteresis in the flow regimes and by tracking the local phase
flow rates it is easy to see how in practice flows may differ greatly from those predicted
by classical steady state methods. It is here that transient two phase flow simulators offer
considerable promise if interface tracking methods can be relied upon. However, one
must be caut ious about the application to real systems. It can be envisaged that, in a long
hilly terrain pipeline, a section may exist close to the slug/stratified boundary. A small
perturbation in the pressure or flow rate may be all that is required to trigger a slug and
change the characteristics of the whole pipeline.
In some cases it has been observed that the exit of a slug in a pipeline apparently triggers
the formation of another slug. Slugs occur regularly with one, two, or three in a line at
one time. This is due in part to the time taken for the liquid level to reestablish, and partly
due to the depacking effect as the slug exits, removing the high pressure drop over the
slug body. This causes an increase in the gas velocity which can trigger another slug.
This is illustrated in Figure 8.1-1 which shows the passage of slugs throughout a 400 m
long, 8-inch diameter, horizontal air/water test facility. The Y-axis shows the
identification number for 30 Light Emitting Diode (LED) slug detector probes, whic h are
located along the top of the pipeline, and give a binary on output when water wet and a
off signal when in air. As the number of slugs in the line increases the influence of
each slug exiting diminishes and slug formation becomes less orderly. The topography of
the pipeline can exaggerate this effect if uphill sections are present at the outlet. This can
increase the pressure fluctuations due to hydrostatic effects, and can lead to severe flow
rate fluctuations such as demonstrated by severe slugging. The random type of slugging
with considerable decay and coalescence is demonstrated in Figure 8.8-2 .
H-0806.35
8-39
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
8-40
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-41
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
In some cases the bubble length is greater than the pipeline length. Here it can take some
time for the liquid level to be reestablished before the next slug is generated. Slugs
generated in this way can often have the longest lengths, as the equilibrium liquid content
in the pipeline can be high. The slug grows as it sweeps up the liquid in front of it,
leaving behind a thin film which may take a considerable time to reestablish. Such
phenomena ma y be well predicted by PLAC or OLGA when the slug tracking model has
been implemented. Otherwise, it may be reasonable to establish a simple maximum slug
length model to be use with steady state simulators. Some have observed this method to
give good agreement with the slugs generated in pipelines under these conditions (e.g.,
BPs Endicott pipeline).
The extrapolation of existing design methods to predict slug flow in hilly terrain pipelines
may be risky, and it is here that dynamic simulation may be of use. This is an area where
validation work is underway. PLAC has been compared with other Sunbury experiments
performed with a pipeline dip configuration on the 6- inch air/water rig, the rig set-up is
shown in Figure 8.8-3. In these tests a known quantity of water was poured into the dip
and allowed to settle. The air flow was then turned on at a constant rate and the flow
behavior observed. The tests were repeated for a range of initial liquid quantities and air
flow rates. At low gas flow rates the slugs are generated at the dip where the liquid depth
is the greatest and leads to a higher local gas velocity which is sufficient to produce
waves that grow into slugs. This is illustrated in Figure 8.8-4. At higher gas velocities
the slug generation point moves downstream of the dip and the slug generation is more
random. This is seen from the densitometer traces illustrated in Figure 8.8-5. The flow
pattern map generated by the matrix of liquid volumes and gas velocities is shown in
Figure 8.8-6, whe re the solid line shows the limiting gas velocity at which liquid is
removed from the dip. At this point the liquid is drained and measured and gives an
indication of the equilibrium liquid holdup. This is compared with a terrain induced slug
flow model, where the agreement is seen to be generally good.
H-0806.35
8-42
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-43
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 8.8-6: Pipeline dip experiments-flow pattern map and equilibrium holdup
Initial comparisons with PLAC have been encouraging as illustrated in Figure 8.8-7,
which shows that PLAC generally predicts the gas velocity at which liquid is removed
from the system and the equilibrium holdup, however PLAC tends to underestimate the
holdup at the higher gas velocities. Figure 8.8-8 shows that the predicted slug frequency
is generally lower than that measured. However, the results are still in reasonable
agreement. This is because PLAC simulates waves rather than slugs, Figure 8.8 -9 shows
the oscillations produced by PLAC for case of 317 liters of water with an air velocity of
0.5 m/s.
H-0806.35
8-44
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 8.8-7: Comparison between PLAC and measured liquid removal limit
H-0806.35
8-45
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 8.8-9: PLAC holdup oscillations for 317 l of water and 0.5 m/s air velocity
H-0806.35
8-46
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
OLGA is also reported to yield good comparison to experiment, though data supporting
this is not generally available; generally restricted to member companies of the OLGA
club; the group of companies actively supporting its development.
In practice oil and gas pipelines will usually reach or attain equilibrium first at the start of
the pipeline, hence slugs can be initiated near the pipeline inlet. It is possible for inlet
flow rate fluctuations to initiate slug formation. Such inlet flow rate oscillations may
result from slugging wells, gas- lift operations, and test well cha nges for example, and
should be suspected as possible causes for flow rate perturbations which generate slugs in
otherwise stratified flowing conditions. When using transient flow simulators it is
worthwhile to consider the effects of practical flow rate perturbations on the result.
8.8.2
H-0806.35
8-47
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-48
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
8-49
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-50
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
slug is produced into the separator it may accelerate due to the reducing gravitational
resistance and the reduced frictional length (Figure 8.8-12). This can cause large velocity
increases which can impact on the process plant control, and can also give rise to large
loads on the vessel internals. This is discussed more fully in Section 17 of this Guideline.
Process plant dynamics can be simulated by a number of codes. BPX Sunbury presently
uses the SPEEDUP code which has also been interfaced with a simple slug hydraulic
model to enable the simulation of slug effects on process plant to be investigated. The
slug size is an input to the simulation. However, it is possible to see the effect of the slug
catcher pressure on the passage of the slug. Classical control schemes generally reduce
the gas compressor speed if the separator gas flow is reduced, hence when the slug is
produced the compressor is decelerating. A better solution can be to increase the
compressor speed when the gas flow drops off, hence reducing the separator pressure and
sucking the slug up the riser. This has the effect of reducing the gas starvation period and
also means that the compressor is accelerating when the gas surge occurs.
A recent study using SPEEDUP to investigate control strategies to limit the impact of
slug dynamics on the process plant has shown that gas outlet flow control can reduce the
peak in the gas surge. As the slug is received the gas flow rate reduces. By opening the
gas outlet control valve, the pressure in the slug catcher is reduced, hence sucking the
slug in and allowing some scope for increase in the slugcatcher pressure to absorb some
of the gas flow rate surge.
8.8.3
H-0806.35
8-51
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
back pressure due to the hydrostatic difference in the levels. Figure 8.8-14 shows the
topography of the test rig as modeled by PLAC whereas Figure 8.8.-15 and Figure 8.8-16
show the PLAC simulations of the liquid draining and inlet pressure variation.
H-0806.35
8-52
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-53
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 8.8-16: PLAC simulation of back pressure due to liquid settling in dips
H-0806.35
8-54
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
While it is seen that PLAC can simulate the liquid settling behavior, the comparison with
the experimental results is not always good. This is because PLAC does not model slugs
properly and hence the holdup in the uphill sections is sometimes underestimated, hence
the initial condition is not always correct.
The highest back pressures can be generated when the flowline is restarted as it is
possible for the liquid in the dips to be displaced downstream before the gas is able to
penetrate the dip, and this increases the difference in the hydrostatic levels. Once the gas
penetrates, the pressure is reduced as the gas lifts the liquid out of the dip. One can
imagine that the start-up pressure can be large if all the liquid in the dips is displaced
simultaneously.
A simple method has been established to estimate the pressure rise during the start-up of
hilly terrain flowlines. For long hilly flowlines, the total uphill elevation changes may be
very large and hence the total theoretical hydrostatic head may lead to very conservative
start-up pressures. It is therefore recommended that the flowlines are started by unloading
the sections closest to the facilities first. The pressure can be estimated by summing the
uphill elevation changes and calculating the hydrostatic head for the section being started.
In practice experience with the start-up of Cusiana T pad indicates that the actual
hydrostatic pressure generated was around 2/3 of the maximum theoretical value,
however this is very dependant on the particular system.
If the analysis of start- up pressures is required to check the design pressure of the
flowline, then the topography should be investigated to determine if there are low points
where the pressure can be higher than at the well pad. This is the case with the T-Q
flowline where the maximum pressures are generated in the valley just downstream of the
wellpad.
8.8.4
Pipeline-riser interactions
A special topographical effect is the interaction between a pipeline and a vertical
platform riser. This can give rise to a phenomenon called severe slugging if the pipeline
inclination is downhill at the base of the riser and if flow rates are low enough (stratified
flow). Severe slugging is discussed in Section 17 and will not be described again here.
However, it is useful to demonstrate that dynamic simulation can be employed to
accurately predict the size of severe slugs and the cycle time, enabling topside process
plant to be sized.
H-0806.35
8-55
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
One of the validation cases for the PLAC code was a comparison with severe slugging
experiments carried-out by BP at Sunbury. The test rig consisted of a 45m flowline
inclined downhill at 2 followed by a 15m vertical riser. The rig operated at near
atmospheric conditions and the test fluids were air and water. A schematic of the rig is
shown in Figure 8.8-17 where the pipeline inner diameter is 2 inches. Figure 8.8.-18 and
Figure 8.8-19 show the measured slug size and frequency for a range of fluid superficial
velocities.
H-0806.35
8-56
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 8.8-20 shows the liquid holdup profiles predicted by PLAC for gas and liquid superficial
velocities of 0.44 m/s and 0.43 m/s respectively. The figure shows the liquid build-up and
blowdown.
H-0806.35
8-57
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-58
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 8.8-21: Figure 7.34 PLAC prediction of outlet liquid mass flow rate.
A similar type of severe slugging can also be observed when flexible risers are employed,
particularly if the lazy s configuration is adopted. The analysis of the flows in these
curved configurations is complicated by the possible bi-directional motion of the flow,
and the interaction with the complex geometry. Here transient simulators show promise
in being able to predict the flows, whereas the scale- up of laboratory simulations to
practical situations is uncertain..
The relationship below can be used to estimate the critical liquid velocity at which severe
slugging occurs in a pipeline-riser system. If the liquid superficial velocity is above this
value severe slugging is unlikely to occur.
Vsl = Vsg Psep / [L(l - g) g (1 - Hl) Sin ]
where:
Psep = pressure at flowline outlet (N/m 2)
H-0806.35
8-59
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-60
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
It is hence seen that the flow split at a simple tee may be very different from equal, and
hence the flow rates in the individual legs will be unequa l. However, since in a loop the
lines rejoin downstream, the number of possible solutions is reduced to those flow splits
that give rise to the same overall pressure drop in both pipelines.
Transient flows can result from the operation of looped pipelines since it is possible for a
number of solutions to exist for the flow split. For example, high flow in one leg causing
a high frictional pressure drop may be balanced by low flow in the other with a high
hydrostatic component and high holdup. Small inlet flow rate perturbations at the tee, or
changes in operating conditions, may cause the flowing conditions in each leg to change,
with the possible removal of excess liquid in the form of a slug. In the extreme it is
possible in pipelines with a large uphill elevation for a manometer effect to exist. Here
the flowing pressure in one leg is balanced by a static column in the other. A change in
operating conditions can cause the static liquid to be swept-out. This type of phenomena
has been known to give rise to operational difficulties with gas transmission pipelines
laid over hilly terrain.
Problems with parallel loops are well documented in the power generation industry,
where instabilities can occur in a bank of boiler tubes, for example. The total flow into
and out of the headers is normally constant, as is the pressure drop. However, flows in
the individual tubes can vary considerably. In such situations the analysis is complicated
by the addition of heat transfer effects and steam generation, which can give rise to
multiple solutions for the location of the boiling front. So called parallel channel
instability in the boiler tubes can cause corrosion problems or dry-out leading to
premature failure. Experience with such situations has shown that adding flow restrictors
to the inlet header has a stabilizing effect, whereas throttling the outlet can exacerbate the
problem.
The method proposed to analyze possible transient problems caused by operating looped
pipelines is along the lines outlined by Gregory and Fogarasi. This involves calculating
the two-phase pressure drop in each leg for a range of liquid and gas flow splits. A
graphical solution is employed to determine the conditions where the pressure drops in
each leg are equal, and hence the possible operating regimes of the loop. Allowance
should then be made for flow excursions between the parallel lines and the possible
displacement of liquid.
As an example consider the schematic shown in Figure 8.8-22.
H-0806.35
8-61
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
8-62
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 8.8-23 shows the locus of the possible solutions for a equal pipeline loop and
indicates that the 50/50 solution (a = 0.5) gives rise to the maximum downstream
pressure, and hence minimum pressure drop, however this is not always the case. For
other pressure drops there are two flow split solutions and hence it could be possible for
the flows to oscillate between the two solutions.
Gregory et al. suggests that the existence of a manometer leg effect can be checked if the
net elevation increase over the loop is such that the maximum possible hydrostatic head
that could result from the liquid phase static in the pipeline is greater than the pressure
H-0806.35
8-63
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
drop that will result from the total throughput of gas and liquid flowing in one of the
parallel lines. This may not be wholly true as a symmetrical hill would have a net
elevation change of zero. However the potential exists for the uphill section to contain
liquid and provide a manometer effect. It is hence recommended that the manometer
effect check be carried out using the sum of the uphill elevation changes. If one of the
legs contains static liquid it is obvious that the loop has in fact not increased the capacity
of the system as the flow only occurs in one leg.
In the above example the pipes in the loop are identical and hence the locus of possible
solutions is symmetrical. This is not the case when the pipes are different as illustrated in
Figure 8.8-24 for a 10- inch and 12- inch loop from Gregorys paper. Here the locus of the
possible solutions is not symmetrical and the minimum pressure drop occurs with a liquid
split of 65/35.
H-0806.35
8-64
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-65
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
3. Although the outlet flow rate from the loop may be steady at one condition, a small
change in the operating conditions can lead to transient liquid sweep-out as the new
operating point stabilizes.
4. A check should be made for manometer effects based on the sum of the uphill
elevation changes, although this may not give the maximum holdup change as a pipe
full flowing case may also exist. If manometer legs occur the loop will not increase
the flowing capacity.
5. Static liquid legs may have implications for corrosion.
8.9
Transient Modeling
In practice the worst transients can be deduced from relatively few simple calculations:
(a) Small diameter tie- ins from minor fields connected to long, large diameter trunklines
do not usually contain enough liquid, even at low flow rates to significantly contribute to
a transient.
(b) For a given total flow in a network, the highest holdup will generally be for that
combination of flows which has the lowest gas flow rate passing through the greatest
length of the largest diameter pipeline.
(c) For low CGR fluids, the maximum period of operation at low gas flow rate should be
considered and the amount of liquid deposited in the pipeline during the period
calculated. This figure may prove less than the steady-state liquid holdup at that flow
rate and if so should be used instead of the steady-state figure for determining the worst
case transients.
Having identified transient flow situations likely to cause problems, a more detailed
analysis can then be performed using a transient modeling code such as PLAC or OLGA.
8.10
H-0806.35
8-66
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
whereas if the changes are made slowly, the liquid can be gradually swept out within the
slug catcher capacity.
It is also interesting to note in this figure the wide variation in the predicted holdup
obtained from using the various methods. This may not be too much of a problem when
calculating the amount of liquid swept out during a transient because it is the difference
in the holdup that it most important. In this example the Eaton holdup correlation is
expected to give reasonable answers. However, one should be wary of using the
correlation approach rather than mechanistic models since they do not usually predict the
steep rise in holdup as the velocity and interfacial friction reduces hence a gross
underestimate can result when starting from low flow rates.
Figure 8.10-2 shows the liquid flow rate at the outlet of the Marlin gas condensate
pipeline during an increase in the flow rate from 155 MMSCFD to 258 MMSCFD.
During the test the gas rate was held constant at 155 MMSCFD for 52 hours in order to
reach equilibrium conditions. The rate was then increased to 258 MMSCFD in a period
H-0806.35
8-67
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
of one hour and held constant for a further 26 hours to obtain equilibrium conditions
again. It is seen that during the transient the outlet liquid flow rate is cons iderably higher
than the final equilibrium value.
Figures 8.10-3 to 8.10-5 are provided to illustrate the potential accuracy of transient
codes on the Marlin rate change data. This data has often been used as a test case as it is
one of the few transient field data sources available in the open literature.
H-0806.35
8-68
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-69
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
8-70
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Pigging gas condensate pipelines can also result in large slugs. However, in many cases
it is not feasible to design slug catchers of a sufficient size to handle the equilibrium slug
produced by running pigs at low throughputs. In this case pigs must be run frequently to
prevent the liquid holdup reaching equilibrium, which can incur high operating costs.
Sometimes gas rates can be reduced during the pig arrival to allow the produced liquids
to be processed, but in others the gas flow rate may be determined by the consumer, and
the pigging operation carried out at the prevailing gas rate. In some situations pigs are
only required on an infrequent basis for corrosion control or inspection, in which case the
pipeline liquid content may be high and a procedure must be put in place to handle the
liquid swept-out by the pig. One way of doing this is to stop the pig offshore and walk
it into the slug catcher at a rate compatible with the liquid processing capacity. This
approach was successfully carried out at the restart of pigging operations on the
Amethyst pipeline where a liquid vo lume of over four times the slug catcher capacity was
allowed to accumulate in the sealine when the offshore pig launcher failed. In some
instances it is possible to reduce the liquid content of the pipeline prior to pigging by
controlled rate increases to remove liquid. In other cases pigging is not possible, and flow
rate changes must be controlled to prevent overfilling the downstream plant.
With these factors in mind it is seen that for gas condensate systems at least, it is often
the transient slug that determines the slug catcher volume. For oil and gas pipelines
pigging may be required frequently for wax control etc, where the lines have reached
equilibrium, hence this may determine the size of the slugcatcher. However for some
developments, particularly subsea, pigging is required less frequently and can be
accomplished with some operating ingenuity. Here it is often the longest normal slug or
the transient rate change liquid sweep out that determines the required slug catcher surge
volume. The next section outlines a simple calculation method for estimating the liquid
outflow profile due to rate change transients.
8.10.2 Simple analysis methods for flow rate increases and reductions
The first step in considering the required surge volume for transient flow rate increases
can be made by considering the equilibrium holdup/flow rate profile.
H-0806.35
8-71
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
8-72
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The holdup profile shows a rapid rise in the liquid content at gas rates below 200
MMSCFD, hence operating in this region could cause problems even without pigging, as
relatively small gas rate perturbations can cause large quantities of liquid to be swept
from the pipeline. For example a 10 percent increase in the gas rate from 200 to 220
MMSCFD could remove 4750 BBL of liquid and, possibly, swamp the slug catcher.
Shortly after the start-up of the Amethyst pipeline system the pig launcher failed and it
was decided to investigate the consequences of cont inuing production, and allowing the
pipeline liquids to build up to equilibrium values. A decision was taken to limit the
minimum gas flow rate to 250 MMSCFD, and hence to avoid possible uncontrolled
sweep-out due to inherent flow rate fluctuations. The pipeline was operated without
pigging over the winter, where it took several months to obtain an equilibrium holdup,
which was estimated to be in the region of 18000 BBL from a mass balance. This is
within 17 percent of the value predicted by the old segregated flow mechanistic model in
MULTIFLO, but is 213 percent higher than the value predicted by the Eaton correlation.
Some of the simple transient analysis outlined below was used to investigate how to
resume pigging operations.
A simple approach to sizing a vessel to handle the liquid produced by a gas rate increase
would be to consider the change in the equilibrium holdup and ignore the effect of the
liquid pump-out rate. For example if the gas rate were increased from 250 MMSCFD to
350 MMSCFD the equilibrium liquid removed would be 15373 - 8018 = 7355 BBL.
Hence the gas rate could be increased in two steps from 250 MMSCFD to 300 MMSCFD
which would remove 3800 BBL and then 300 MMSCFD to 350 MMSCFD which would
remove 3555 BBL. A 7355 BBL surge volume would be required if the increase were
made in one step without taking account of the liquid pump-out rate. The maximum gas
flow rate available is 350 MMSCFD which gives a pipeline inventory of 8018 BBL, and
hence it is still necessary to walk in the first pig.
We will use the Amethyst case to illustrate a simple way of determining the effect of the
pump-out rate on the required surge volume. Consider the case of an increase in the gas
rate from 250 to 350 MMSCFD.
H-0806.35
8-73
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Next calculate the initial and final liquid production rates from:
Liquid rate = liquid loading x gas flow rate
hence:
Initial liquid flow rate = 250 MMSCFD x 7.04 BBL/MMSCF = 1760 BPD
Final liquid flow rate = 350 MMSCFD x 7.04 BBL/MMSCF = 2464 BPD
Calculate the duration of the transition time for the transient, which is the length of time
over which the high flow rate occurs. If it is assumed that all the liquid in the line
accelerates to the equilibrium liquid velocity corresponding to the final gas rate, then the
transition time is the same as the residence time at the final rate, i.e.:
Transition time = (final holdup / final flow rate) = (8018 / 2464) = 3.25 days
The transition flow rate is the sum of the final flow rate and the increase due to the rate
change and is given by:
Transition flow rate = Final flow rate + (holdup change / transition time)
= 2464 + 7355/3.25 = 4727 BPD
It is seen that based on this method the surge can easily be handled by using the 6000
BPD pump-out capacity of the Easington terminal. Figure 8.10-7 shows the predicted
liquid outflow profile.
H-0806.35
8-74
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
This approach can be extended to investigate the trade-off between the required slug
catcher surge volume and the pump-out rate using the relation below:
Surge volume = transition time x (flow rate in - flow rate out)
= Tt x (Qin - Qout)
If the pump-out rate is fixed at the final equilibrium value the required surge volume is:
Vs = 3.25 ( 4727 - 2464 ) = 7355 BBL
i.e. the change in equilibrium holdup.
If the pump-out rate is 4727 BPD then this method shows that no surge volume is
required. The solution to the equation is the linear relationship shown in Figure 8.10-8. It
can be seen that for a surge volume of 3800 BBL a minimum pump-out rate of 3550 BPD
is required.
H-0806.35
8-75
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The relationship shown in Figure 8.10-8 can be subject to large inaccuracies at the
extremes as the flow rate tends to zero and at the final transition flow rate. The reason
being is as follows; at zero pump-out rate the surge volume is equal to the transition flow
rate multiplied by the transition time, whereas in practice liquid continues to flow into the
vessel at the final equilibrium flow rate, hence the required surge volume becomes
infinite as the pump-out rate goes to zero. When the pump-out rate is equal to the
transition flow rate the above method indicates that no surge volume is required.
However, in practice, the flow rate during the transient is not usually constant, and
typically peaks at the start. Hence, the solution for a surge volume of zero is a pump-out
rate equal to the peak flow rate during the transient.
The same simple method can also be applied to the estimation of the outlet flow rate
profile during a flow rate decrease, as follows: Consider a reduction in the gas flow rate
of the Amethyst pipeline from 350 MMSCFD to 250 MMSCFD.
First calculate the residence time at the final flow rate, this is also assumed to be the
transition time:
Transition time = (final holdup / final flow rate) = (15373 / 1760) = 8.73 days
then:
H-0806.35
8-76
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Transition flow rate = Final flow rate (Holdup change / Transition time)
= 1760 (7355 / 8.73) = 918 BPD
Hence the hand calculation method predicts an initial flow rate of 2464 BPD falling to
918 BPD over a 8.73 day transition period after which the rate increases to 1760 BPD.
This is illustrated in Figure 8.10-9.
H-0806.35
8-77
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
maximum oil production from the subsea wells was expected to be 12 mbd and slug
catchers were required to be sized for the platform reception facilities.
The slugs produced by pigging the pipeline were to be handled by operational procedures
and were not used for sizing the slug catcher. The predicted mean and maximum normal
hydrodynamic slugs were 6 BBL and 16 BBL respectively, and were well within the
feasible slug catcher size, which was in the region 50100 BBL. It was hence necessary
to consider the slugs produced by flowline rate changes which could be quite frequent as
wells are switched to test and as turndowns are accommodated. A worst case was
considered to be start-up from 1 well to full production, giving a flow rate increase from
112 mbd. This was modeled with PLAC by running the simulation at 12 mbd for
13000s to give a steady state. The flow rate was then ramped down to 1 mbd over 60s,
left at this rate for 5000s, then increased to 12 mbd over 60s. The predicted outlet liquid
flow rate profile is shown in Figure 8.10-10 and indicates that the liquid flow rate drops
off when the flow rate is reduced after 13000s. The plot shows a large overshoot when
the rate is increased again at 18000s. The oscillations decay as the final equilibrium
production rate of 12 mbd (equivalent to 17.7 kg/s) is approached. The peak production
rate is 82 kg/s which is equivalent to 51 mbd, and is hence over four times the final
equilibrium production rate.
8-78
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The outlet gas rate is shown in Figure 8.10-11 and indicates that the main gas surge
occurs after the second slug.
The liquid holdup plot in Figure 8.10-12, shows that the initial holdup is around 40
percent in the inclined flowline (cells 118) and around 20 percent in the vertical riser
(cells 1939). During the low flow condition from 13000s the liquid drains from the riser
into the flowline and also builds up at the template end of the flowline. During the rate
increase the first slugs were due to the liquid drained from the riser. The liquid that
drained to the template is smeared out along the flowline to re-establish the equilibrium
holdup.
H-0806.35
8-79
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-80
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 8.10-13: Slug catcher level fluctuations with and without level control
The result is also plotted in Figure 8.10-14 for a pump-out rate of 12 mbd and shows that
the vessel would be swamped, even without low level control. Hence it is not possible to
H-0806.35
8-81
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
use a 50 BBL slug catcher if the liquid pumpout rate remains fixed at the maximum
subsea production rate.
It is possible by iteration to determine the vessel size vs pump-out rate relationship that
just handles the inlet flow transient. This is shown in Figure 8.10-15 and illustrates that if
the pump-out rate is to remain at the normal subsea production rate of 12 mbd, then a 275
BBL slug catcher would be required, but if the spare capacity of the platform process is
utilized to give a pump-out rate of 25 mbd, then the required slug catcher volume is
around 40 BBL.
H-0806.35
8-82
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 8.10-16: MULTIFLO holdup predictions for Pompano 6-inch subsea flowline
It is seen that for an initial rate of 1 mbd the equilibrium holdup is 770 BBL for 0 percent
water cut, reducing to 470 BBL at a final flow rate of 12 mbd.
The residence time is approximated by the equilibrium holdup divided by the liquid flow
rate, hence:
H-0806.35
8-83
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 8.10-17: Approximate outlet liquid flow rate profile and PLAC prediction
The slug catcher volume vs pump-out relationship is hence given by a simple linear
function since the flow rate during the transition is assumed to be constant, therefore:
Surge volume = 0.039 x (19692 - Qout)
H-0806.35
8-84
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
hence:
at 0 mbd,
at 12 mbd,
at 19.692 mbd,
The comparison between the simple method and the rigorous PLAC simulation is shown
in Figure 8.10-18 where it is seen that the agreement is close for pump-out rates similar
to the final equilibrium value. However, there are large discrepancies at the extremes for
the reasons outlined previously. For zero surge volume the pump-out rate predicted by
PLAC is 51 mbd compared to 19.7 mbd assuming a constant transition flow rate, hence
the simple method fails to take account of the liquid distribution in the pipeline and
consequently underestimates the peak outlet flow rates.
Figure 8.10-18: Comparison of slug catcher relationships derived from PLAC and hand
calculations
8-85
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Peterhead at 30 degree C and 25 bara. During shut-down and start-up there is the
possibility of liquid condensation and drop-out which may affect the size of the
downstream plant and operating procedures necessary to control the liquid outflow. The
combination of rate changes and associated cooldown gives rise to a more complicated
transient analysis than previously discussed.
In such a transient the amount of liquid formed is a function of the richness of the input
gas composition, the temperature gradient along the pipeline, Joule-Thompson effects,
transient cool-down, the topography, and pressure drops. Some of the interactions
between these effects can be complicated. For example, the pressure drop can cause
retrograde condensation which can give rise to higher pressure drops due to interphase
friction. This can lead to J-T cooling which produces more liquid dropout, and so on.
Once liquid is produced it may take time to drain into dips, and may produce slugs. On
start-up the liquid can re-evaporate in the pipeline before reaching the outlet.
The OLGA code was used to simulate the effects caused by the shut-down and
consequent cooling of the Miller landline. The start-up was also modeled in order to
study the complex transient multiphase effects taking place throughout the pipeline
system as a result of the introduction of warm gas into the cool environment. This study
formed part of the design stage of the pipeline project and provided a valuable insight
into the sizing requirements for slug-catching equipment at the power station delivery
point.
The transient simulations were performed in three stages:
1. Run to steady state at 215 MMSCFD with an inlet pressure of 30 bara and an inlet
temperature of 37C.
2. Shut-down with the outlet closed and the line allowed to pack up to 35 bara
throughout. The line is subsequently allowed to cool for 48 hours.
3. Start- up the pipeline with inflow and outflow of 60 MMSCFD and remain constant
for 13 hours. The inlet flowing gas temperature was the order of 80C. Finally, the
flow rate is ramped to 215 MMSCFD with an inlet temperature of 37C and the
simulation continued until the outlet temperature exceeds 30C.
The results of the simulation are best described with reference to the pipeline profile and
the relevant cell numbers illustrated in Figure 8.10-19
H-0806.35
8-86
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-87
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-88
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
In the first phase of the start-up the gas flow rate is ramped up to 60 MMSCFD with an
inlet temperature of 80C. This has the effect of reducing the downstream pressure as the
hydraulic gradient is established, and warming the fluid temperature. The warm- up is
seen to be slow as the outlet gas temperature has only risen by a few degrees after 13
hours (Figure 8.10-22). The slow warm- up and flow rate increase produces liquid
sweepout at some of the dips. However the liquid fills subsequent dips and does not exit
the pipeline, hence there is still a net increase in the pipeline liquid content during the
first phase of the start-up (Figure 8.10-23).
H-0806.35
8-89
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure
8.10-23: Pipeline liquid content during first start -up phase
In the second start-up phase the gas rate is increased to 215 MMSCFD with the gas
flowing in at 38C and 35 bara. The further increase in the flow rate results in a higher
pressure gradient and hence a decrease in the outlet pressure. The increase in the cold gas
inflow rate causes the warm front to accelerate through the pipeline (Figure 8.10-24)
causing liquid to be flashed off, and sweeps out the residual liquid from the system
(Figure 8.10-25). As a result the total liquid content of the system drops sharply removing
around 110 m3 (690 BBL) of liquid from the pipeline (Figure 8.10-26), however around
30 percent of this liquid is evaporated and the remainder flows to the slug catcher.
H-0806.35
8-90
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-91
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 8.10-26: Pipeline liquid content during second start -up phase
8.10.6 PLAC analysis of severe slugging in a catenary riser
A comparison has been made between PLAC and some experimental results of severe
slugging in a Catenary riser. The experimental data was taken by BHRA in 1990
(Reference 1) and involves holdup, pressure, and velocity measurements in a 2-inch
diameter, 108 ft high catenary riser model using air and water as the test fluids. A 200 ft
length of 2 downhill inclined line was used before the riser and an air buffer vessel of
0.126 m3 was used in the air supply line to model a larger pipeline length. The test rig is
illustrated in Figure 8.10-27.
PLAC simulations have been performed by XFE for a test condition just in the severe
slugging region corresponding to gas and liquid superficial velocities of 2.3 m/s and 0.33
m/s respectively. The measured pressure at various points in the riser are shown in
Figure 8.10-28 where the top traces are for the transducers in the inclined flowline and at
the base of the riser. The experimental severe slugging cycle time is 183 seconds.
The test rig was modeled in PLAC as a TEE component with 86 cells, in which a 31m
horizontal section of pipe was used to simulate the air buffer vessel. The liquid is
introduced into the side arm of the TEE, which is located at the end of the horizontal
section. The topography is shown in Figure 8.10-29. The PLAC simulations begin with
H-0806.35
8-92
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
an empty pipe and hence some time is required for the severe slugging cycle to be
established. Figure 8.10-30 shows the pressure in the inclined flowline, which is in good
agreement with the measured data of Figure 8.10-28.
H-0806.35
8-93
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-94
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Cycle time
Slug build- up time
Slug production time
Bubble penetration time
Gas blowdown time
Slug length
Maximum pressure at riser base
Slug tail exit velocity
Experimental
measurements
183 s
124 s
24 s
14 s
21 s
59 m
47.5 psig
3.5 m/s
PLAC
predictions
172 s
127 s
11 s
9s
25 s
57 m
48 psig
4.5 m/s
From the above comparison it is seen that in most cases the PLAC predictions are in good
agreement with the measured values apart from the slug production time. This time is
however rela tively short and is difficult to accurately determine from the plots. The
experimental conditions were not tabulated by BHRA and hence the estimation of the
H-0806.35
8-95
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
flowing velocities could also be subject to some error. However, the predictions indicate
that the flowing conditions in the riser are close to the boundary for true severe slugging
as the riser is only full of liquid for a short period of time before the gas pressure is
sufficient to eject the slug. Figure 8.10-34 shows the test case point on the experimental
flow pattern map and confirms that the PLAC predictions are qualitatively correct.
H-0806.35
8-96
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-97
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-98
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-99
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The table below shows the calculated gas-oil ratios for a 70 mbd oil flow rate with a
producing gas-oil ratio of 2000 scf/stbo. At the inlet conditions to the loop of 600 psia
and 175F the calculated solution gas-oil ratio is 180.81 scf/stbo. a is the fraction of the
inlet liquid flow rate in line A and is the fraction of the inlet free gas flow rate in line A.
Table 8-3: Effective GORs for use in two -phase simulations
Liquid fraction ?
0.1
0.2
0.4
0.6
0.8
0.9
1.0
0.0
181
181
181
181
181
181
181
0.1
2000
1090
636
484
408
383
363
0.2
3819
2000
1090
787
636
585
545
0.4
7458
3819
2000
1394
1090
989
908
0.6
11096
5638
2910
2000
1545
1394
1272
0.8
14734
7478
3819
2606
2000
1798
1636
0.9
16554
8367
4274
2910
2227
2000
1818
1.0
18373
9277
4729
3212
2455
2202
2000
Qoa
14
28
42
56
63
70
The MULTIFLO predicted pressure drops for each gas and liquid flow rate combination
are shown in the table below where the pressure drops are calculated for a fixed value of
at each value of a and hence GOR and oil flow rate Qoa.
Table 8-4: Overall two -phase pressure drops for line A
0.1
0.2
Liquid fraction
0.4
0.6
0.8
0.0
0.1
0.2
163
118
91
162
130
101
163
123
95
164
109
96
167
122
112
169
129
120
171
135
128
0.4
0.6
68
65
83
78
75
80
94
104
114
127
124
138
134
149
0.8
0.9
1.0
68
71
75
82
86
91
92
99
108
118
127
137
144
154
165
157
168
179
170
181
193
H-0806.35
8-100
0.9
1.0
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The data can be plotted as pressure drop vs a for each value of . If the line sizes or
topographies of the loop are different the calculations must also be repeated for line B.
However, if both lines are the same the results will be symmetrical i.e. a = 0.9 and = 0.9
for line A corresponds to a = 0.1 and = 0.1 for line B. The result for the case considered
is shown in Figures 8.10-36 and 8.10-37, where the locus of possible solutions is shown
in Figure 8.10-37.
H-0806.35
8-101
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-102
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
For the example above the sum of the uphill elevations is 504 ft, which gives rise to a
hydrostatic head of 163 psi. Figure 8.10-38 indicates that manometer effects are possible
when the inlet flow rate is reduced below around 60 mbd since the flowing pressure drop
could be balanced by a static column of liquid in the other leg.
Figure 8.10-38: Pressure drop characteristic for total flow in the one leg
H-0806.35
8-103
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
It was required to estimate the loads produced by the pig-slug impacting on the PCV if it
closed during pig reception, and hence to determine the need to set the valve fully open
during pigging. The slug catcher is located 10m above ground level and it was required to
assess the structural loads that may be generated by the slug dynamics.
A simple dynamic model of the slug reception process indicated that the slug velocity
increases from a initial value of 7m/s to a final velocity of 21m/s as the tail passes the
PCV and the slug is discharged into the slug catcher. This acceleration is due partly to the
hydrostatic head loss and the reducing frictional length of the slug as it is produced.
Initial simulations using FLOWMASTER were based on the surge pressures generated
by closing the PCV during slug reception. A linear valve closure rate was assumed and
the effect of the valve closure time investigated. The results of a single phase surge
analysis are shown in Figure 8.10-40 indicating that surge pressure should not exceed 95
bar, which is well below the maximum allowed pressure of 150 bar. The valve closure
times of around 30s are too long to generate significant unequilibriated loads in the short
piping runs. What was of more concern was the impact of the slug front on a partially
open control valve, which is more difficult to assess.
H-0806.35
8-104
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
One of the major unknowns is the time over which the slug front impact occurs. The
worst case is to assume a vertical slug front and hence an instantaneous impact. The
results of using the primer component in FLOWMASTER to simulate this is shown in
Figure 8.10-41 for locations F, G, and H. The slug was in this case assumed to be
travelling at 6 m/s through a 20 percent open control valve. An all liquid slug was
assumed.
H-0806.35
8-105
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-106
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
The reaction loads due to the unhindered slug passage are more easily calculated. These
are due to the change in the momentum between the liquid slug and the gas and are
related to the slug front and tail velocities. Ignoring elevation effects the slug front
velocity is constant and hence, the coming on load is fixed. However, the slug tail
accelerates during slug reception, and hence the largest loads are the coming off loads
produced by the passage of the slug tail. These loads are shown in Figure 8.10-43 and
indicates that the maximum load is of the same order as that produced by a 50ms impact
surge.
H-0806.35
8-107
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Information Sources
Bevan, D.J. "Correlation more accurately predicts two-phase pipeline holdup" Oil & Gas
Journal, p81 - 88, 20th Apr. 1992.
Cunliffe, R.S. "Condensate flow in wet- gas lines can be predicted" Oil & gas Journal,
p100 - 108, 30th Oct. 1978.
Eaton, B.A. Andrews, D.E., Knowles, C.R., Silberberg, I.H. & Brown, K.E. "The
Prediction of Flow Patterns, Liquid Holdup, and Pressure Losses Occurring during
Continuous Two-phase Flow in Horizontal Pipelines" Journal of Petroleum. Tech., p815
- 828, June 1967.
Gregory, G.A. "Multiphase Flow in Pipes - Prediction of Delivered Liquid Flow Rate for
Changes in Total Flow Rate in a Two Phase Pipeline" Notes for Course Presented by
Neotechnology Consultants Ltd., Calgary, Alberta, Canada, 1991.
Gregory, G.A and Fogarasi, M Estimation of pressure drop in two-phase oil- gas looped
pipeline systems, Pipeline Technology, March-April 1982, pp75-81.
H-0806.35
8-108
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
8-109
1-Dec-00
9.
HYDRATES
9.1
Introduction
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Natural gas hydrates are crystals formed by water with natural gases and associated
liquids, in a ratio of 85 mole % water to 15% hydrocarbons. The hydrocarbons are
encaged in ice- like solids, which do not flow, but rapidly grow and agglomerate to
sizes, which can block flow lines. Hydrates can form anywhere and anytime that
hydrocarbons and water are present at the right temperature and pressure, such as in
wells, flow lines, or valves and meter discharges. Appendix A gives hydrate crystal
details at the molecular level, along with similarities and differences from ice.
The low temperatures and high pressures of the deepwater environment cause hydrate
formation, as a function of gas and water composition. In a pipeline, hydrate masses
usually form at the hydrocarbon-water interface, and accumulate as flow pushes them
downstream. The resulting porous hydrate plugs have the unusual ability to transmit
some degree of gas pressure, while they act as a flow hindrance. Both gas and liquid
can frequently be transmitted through the plug; however, lower viscosity and surface
tension favors the flow of gas. Depressurization of pipelines is the principal offshore
tool for hydrate plug removal; depressurization sometimes prevents normal production
for weeks.
This handbook was written to provide the offshore facilities/design engineer with
practical answers to the following four questions:
What are the safety problems associated with hydrates? (Section 9.2)
What are the best methods to prevent hydrates? (Section 9.3)
How are hydrate plugs best removed? (Section 9.4)
What are the economics for prevention and remediation? (Section 9.5)
Field case studies, pictures, diagrams, and example calculations are the basis for this
handbook. Less pressing questions regarding hydrate structures, plug formation
mechanism, etc. are considered as background material in Appendix A. A computer
program disk and Users Guide (Appendix B) are provided to enable prediction of
hydrate conditions. Appendix C is a compilation of Case Studies not in the handbook
body. A Russian hydrate perspective is presented in Makogons (1981, 1997) books.
An in-depth, theoretical hydrate treatment is given by Sloan (1998).
H-0806.35
9-1
1-Dec-00
9.2
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Hydrate densities are like that of ice; a dislodged hydrate plug can be a projectile
with high velocities. In the 1997 DeepStar Wyoming field tests, plugs ranged
from 25-200 ft with velocities between 60-270 ft/s. Such velocities and masses
provide enough momentum to cause two types of failure at a pipeline restriction
(orifice), obstruction (flange or valve), or sharp change in direction (bend, elbow,
or tee) as shown in Figure 9.2.1-1. First, hydrate impact can fracture pipe, and
second, extreme compression of gas can cause pipe rupture downstream of the
hydrate path.
2.
Hydrates can form either single or multiple plugs, with no method to predict which
will occur. High differential pressures can be trapped between plugs, even when
the discharge ends of plugs are depressurized.
3.
Hydrates contain as much as 180 volumes (STP) of gas per volume of hydrate.
When hydrate plugs are dissociated by heating, any confinement causes rapid gas
pressure increases. However, hydrate plug heating is not an offshore option due to
the difficulty of locating the plug and economics of heating a submerged pipeline.
Field engineers discuss the hail-on-a-tin-roof sounds when small hydrate particles
hit a pipe wall. Such small, mobile particles can accumulate to large masses
occupying a considerable volume, often filling the pipeline to tens or hundreds of feet
in length. Attempts to blow the plug out of the line by increasing upstream pressure
(see Rule-of-Thumb 18) will result in additional hydrate formation and perhaps
pipeline rupture.
When a plug is depressurized using a high differential pressure, the dislodged plug can
be a dangerous projectile which can cause pipeline damage, as the below three case
studies (from Mobils Kent and Coolen, 1992) indicate.
9.2.1
H-0806.35
9-2
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Figure 9.2.1-1
9.2.3
H-0806.35
9-3
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-4
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-5
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
In the above three case studies several common equipment circumstances existed. The
systems:
1.
2.
3.
4.
5.
The Chevron Canada Resources Hydrate Handling Guidelines (1992) suggest that the
danger of line failure due to hydrate plug(s) is more prevalent when:
long lengths of pressurized gas are trapped upstream,
low downstream pressures provide less cus hion between a plug and restriction, and
restrictions/bends exist downstream of the plug.
9.2.4
9.2.5
H-0806.35
9-6
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-7
1-Dec-00
9.3
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Free water and natural gas are needed. Gas molecules ranging in size from
methane to butane are typical hydrate components, including CO2 , N2 , and H2 S.
The water in hydrates can come from free water produced from the reservoir, or
from water condensed by cooling the gas phase. Usually the pipeline residence
time is insufficient for hydrates to form either from water vaporized into the gas,
or from gas dissolved in the liquid water.
2.
3.
The above three hydrate requirements lead to four classical thermodynamic prevention
methods:
H-0806.35
1.
Water removal provides the best protection. Free water is removed through
separation, and water dissolved in the gas is removed by drying with tri-ethylene
glycol to obtain water contents less than 7 lb m/MMscf. Water removal processing
is difficult and costly between the wellhead and the platform so other prevention
schemes must be used.
2.
Maintaining high temperatures keeps the system in the hydrate-free region. High
reservoir fluid temperature may be retained through insulation and pipe bundling,
or additional heat may be input via hot fluids or electrical heating, although this is
not economical in many cases.
3.
The system may be decreased below hydrate formation pressure. This leads to the
concept of designing system pressure drops at high temperature points (e.g.
bottom- hole chokes). However, the resulting lower density will decrease pipeline
efficiency.
9-8
1-Dec-00
4.
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Two kinetic means of hydrate inhibition have been added to the thermodynamic
inhibitor list and are being brought into common practice:
1.
Kinetic inhibitors are low molecular weight polymers and small molecules
dissolved in a carrier solvent and injected into the water phase in pipelines. These
inhibitors work by bonding to the hydrate surface and preventing crystal
nucleation and growth for a period longer than the free water residence time in a
pipeline. Water is then removed at a platform or onshore.
2.
The above methods are used individually or jointly for prevention. The prevention
section of this handbook provides a method to use the six above methods to prevent
hydrates in the design of an offshore system.
Hydrates form in offshore systems in two fundamental ways: (a) slow cooling of a
fluid as in a pipeline or (b) rapid cooling caused by depressurization across valves as
on a platform.
Section 9.3.1 provides typical offshore system examples of hydrate formation in a
well, a flowline, and a platform. Offshore design for hydrate thermodynamic
inhibition with slow cooling of a pipeline is the topic of Sections 9.3.2, 9.3.3, 9.3.4,
and 9.3.5. Design practices are provided in Section 9.3.6 fo r hydrate prevention with
rapid cooling across a restriction like a valve. Section 9.3.7 gives procedures for
prevention of hydrates through inhibition and heat management. Section 9.3.8
provides general design guidelines for hydrate prevention in an offshore system.
9.3.1
H-0806.35
9-9
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
prevent hydrate formation, and after the platform export lines have dry gas and
oil/condensate with insufficient water to form hydrates.
In Figure 9.3.1-1, two unusual aspects of the system should be noted: (1) the water
depth is shown as 6,000 feet but it may range to 10,000 feet, and (2) the distance
between the well and the platform may range to 60 miles. Such depths and distances
provide cooling for the pipeline fluids to low temperatures, which are well within the
hydrate stability region.
The system temperature and pressure at the point of hydrate formation must be within
the hydrate stability region, as determined by the methods of Sections 9.3.2 through
9.3.4. The system temperature and pressure enters into the hydrate formation region,
either through a normal cooling process (Example 2 and Figures 9.3.1-4 and 9.3.1-5)
or through a Joule-Thomson process (Section 9.3.6).
A typical plot of the water temperature in the Gulf of Mexico is shown in Figure
9.3.1-2 as a function of water depth. The plot shows a high temperature of 70 o F (or
more) occurs for the first 250 feet of depth. However, when the depth exceeds 3,000
feet the bottom water temperature is very uniform at about 40 o F, no matter how high
the temperature is at the air-water surface. This remarkably uniform water
temperature at depths greater than 3,000 feet occurs in almost all of the earths oceans,
(caused by the water density inversion) except in a few cases with cold subsea
currents.
The ocean acts as a heat sink for any gas or oil produced so that, without insulation or
other heat control methods, any flowline fluid cools to within a few degrees of 40 o F,
no further than a few miles of the wellhead. The rate of cooling with length is a
function of the initial reservoir temperature, the flow rate, the pipeline diameter, and
other fluid flow and heat transfer factors. However, as shown in Section 9.3.2, the
ocean bottom temperature of 40 o F is low enough to cause hydrates to form at any
typical pipeline pressure.
H-0806.35
9-10
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-11
1-Dec-00
9.3.1.1
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-12
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-13
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-14
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Davalath and Barker (1993) provide a comprehensive set of conditions for dealing
with hydrates in deepwater production and testing, including two case studies of
problems (summarized in Appendix C Case Studies C.23 and C.24) and four case
studies of successful hydrate management. Typically methanol injection capability is
provided in the well at two places: (1) at the subsea tree, and (2) downhole several
thousand feet below the seafloor. The injection location and amount of methanol
injection are specified using the procedure indicated in Section 9.3.7.1 on methanol
injection.
In offshore well drilling, frequently a water-based drilling fluid is used that can form
hydrates and plug blow-out preventors, kill lines, etc. when a gas bubble (or kick)
comes into the drilling apparatus. This represents a potentially dangerous situation for
well control. Hydrate formation on drilling is an area of active research with several
joint industrial projects underway. While a brief overview is given here, the reader is
referred to Sloan (1998, Section 8.3.2) for a detailed discussion.
Barker indicated the following rules-of-thumb used by Exxon in considering hydrate
formation with drilling fluids.
Drilling hydrate problems frequently occur, but have only been recognized in recent
years.
When hydrates form solids, they remove water from the mud, leaving a solid barite
plug.
One should not design a well to operate outside the hydrate region only if flow
conditions are maintained. If the well will be in the hydrate formation region at static
conditions, flow will stop at some period and the well operation will be jeopardized.
Several hours may be required for hydrate formation and blockage to occur.
As of October 1988 Exxon used salt at the saturation limit range of 150 to 170 g/l to
prevent hydrate formation.
As general guidelines concerning hydrate formation at various water depths, the
summary given below in Table 3.1-1 by Barker may be used.
H-0806.35
9-15
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Water Depth
(ft.)
<1000
1500
2000
3000
By 1988 Shell had drilled 16 wells in the Gulf of Mexico at water depths between
2,000 and 7,500 feet, using muds with 20 wt% sodium chloride (NaCl) and partially
hydrolyzed polyacrylamide (PHPA). In each well Shell experienced an average of
more than one gas kick per well, which signaled the possibility of hydrate formation.
Only one instance in 2900 ft of water involved the possibility of hyd rate formation,
when Shell experienced difficulty disconnecting the drill stack.
Barker and Gomez (1989) documented two occurrences (see Case Studies C.21 and
C.22 of Appendix C) of hydrate formation in relatively shallow waters off California
and the Gulf of Mexico, where losses in drill times were 70 days and 50 days,
respectively. Recently, the number of hydrate problems have increased dramatically
as drilling has moved to deeper water. In several cases where safety was an issue
(plugged blow out preve ntors, stack connectors, etc.) the well was abandoned. Much
remains to be done in this area.
Downstream of the well and choke, the fluid flows through a pipeline of considerable
length before reaching the platform. Example 2 represents flow conditions in the
pipeline.
9.3.1.2
H-0806.35
9-16
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
(e.g. 7 miles) the flowing stream retains a high temperature from the hot reservoir gas
at the pipeline entrance. The ocean cools the system, and at about 9 miles a unit mass
of flowing gas and associated water enters the hydrate region (shaded region to the left
of the line marked 0% MeOH), remaining in the uninhibited hydrate area until mile
45. Such a distance may represent several days of residence time for the water phase,
so that hydrates would undoubtedly form, were not inhibition steps taken.
In Figure 9.3.1-4, by mile 25 the temperature of the pipeline system is within a few
degrees of the ocean floor temperature, so that approximately 23 wt% methanol is
required in the free water phase to prevent hydrate formation and subsequent pipeline
blockage. Methanol injection facilities are not available at the needed point along the
pipeline. Instead methanol is injected into the pipeline at the subsea wellhead. In the
case of the pipeline shown in Figure 9.3.1-4 methanol is injected at the wellhead so
that in excess of 23 wt% methanol will be present in the free water phase over the
entire pipeline length.
As vaporized methanol flows along the pipeline in Figure 9.3.1-4, it dissolves into any
produced brine or water condensed from the gas. Hydrate inhibition occurs in the free
water, usually at accumulations with some change in geometry (e.g., a bend or
pipeline dip along an ocean floor depression) or some nucleation site (e.g., sand, weld
slag, etc.).
Hydrate inhibition occurs in the aqueous liquid, rather than in the vapor or condensate.
While most of the methanol dissolves in the water phase, a significant amount of
methanol either remains with the vapor or dissolves into any liquid hydrocarbon phase
present as calculated using the methods shown later in this section.
In Figure 9.3.1-4, Notz showed that the gas temperature increases from mile 30 to mile
45 with warmer (shallower) water conditions. From mile 45 to mile 50 however, a
second cooling trend is observed due to a Joule-Thomson gas expansion effect.
Methanol exiting the pipeline in the vapor, aqueous, and condensate phases is usually
not recovered, due to the expense of regeneration.
Todd (1997) provided simulations with a different behavior from the pipeline in
Figure 9.3.1-4. In Todds simulations, typical gas pipeline pressure drops are small
relative to the overall pressure, resulting in an almost constant pressure cooling,
providing a straight, horizontal line between the pipeline end points on a plot like
Figure 9.3.1-5. Pipeline pressure drops are functions of several variables, and
individual systems should be simulated for best results.
H-0806.35
9-17
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
FIGURE 9.3.1-4: OFFSHORE PIPELINE PLOTTED ON HYDRATE FORMATION CURVES (FROM NOTZ 1994)
H-0806.35
9-18
1-Dec-00
9.3.1.3
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
to separate the gas, water, and oil, providing an oil phase which has a very low
vapor pressure, and providing water discharge to the ocean.
4.
to dehydrate the gas to a water content below 7 lb m/MMscf before injection into
the pipeline to shore, and
5.
Note that water separation and gas dehydration are vital for hydrate prevention, so that
even if the system cools into the hydrate pressure-temperature region shown in Figure
9.3.1-5, hydrate formation is prevented due to insufficient water. The export pipeline
gas water content is below its water dew point (9 lb m/MMscf) at the lowest
temperature (39 o F) so free water will not condense from the gas phase.
The oil is stabilized by flow through a series of four separators, operating at 1000 psig,
300 psig, 55 psig, and 2 psig before the export oil pipeline, so an oil pipeline pressure
greater than 15 psia will prevent a gas phase. Hydrate formation is not a significant
problem in the oil export pipeline because relatively few hydrate formers (nitrogen,
methane, ethane, propane, butanes and CO2 ) are present and the water content is low.
The gas from each separator is compressed, cooled, and separated from liquid again
before re-combining the gas with the previous separators gas for injection into the
export gas line. The additional oil obtained after cooling the compressed gas amounts
to about 1.5% of the total oil production.
H-0806.35
9-19
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-20
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
FIGURE 9.3.1-6:TYPICAL OFFSHORE PLATFORM SCHEMATIC (FROM MANNING AND THOMPSON 1991)
H-0806.35
9-21
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-22
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-23
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
In the process shown, 4310 bhp compressors represent the largest cost on the platform,
with capital cost on the order of $800-$1500 (1990 dollars) per installed horsepower.
These compressors are powered by fuel gas, which operates at a low pressure (about
200 psig), usually fed from the inlet gas passing through a control valve with a
substantial pressure reduction.
Pressure reductions after the fuel gas takeoff cause cooling, so that point is very
susceptible to hydrate formation, particularly in winter months. Also instrument gas
lines require similar pressure reductions from a header. Texacos Todd et al. (1996.
pp. 35-42) observe that when fuel and/or instrument gas lines are blocked due to
hydrates, the process frequently shuts down, resulting in pipeline cooling and
significant hydrate blockages in the production line at restart.
Hydrate limits to pressure reductions through restrictions such as valves and orifices is
shown in Section 9.3.6.
9.3.2
H-0806.35
9-24
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Hydrate formation data were averaged for 20 natural gases (from Sloan, 1998, Chapter
6) with an average formation pressure of 181 psia. Of the 20 gases, the lowest
formation pressure was 100 psig for a gas with 7 mole % C3 H8 , while the highest
value was 300 psig for a gas with 1.8 mole % C3 H8 .
Rule-of-Thumb 1 indicates that most offshore pipeline pressures greatly exceed the
hydrate formation condition, indicating:
gas drying and/or inhibition is needed for ocean pipelines with temperatures
approaching 39 o F,
a more accurate estimation procedure should normally be considered, and
hydrate formation pressures are dependent upon the gas composition, and are
particularly sensitive to the amount of propane present.
It should be reiterated here that hydrates can form at temperatures in excess of 39 o F
when the pressure is elevated, as in the case of warmer temperatures in shallower
water. More accurate estimations of hydrate formation conditions over a broad
temperature range are made by the method in the following section.
9.3.3
2.
The intersection of the above two lines determines the pressure and temperature at
which hydrates will form in a pipeline. As we have seen in Exa mple 2 of Section
9.3.1, it is very likely that a long offshore pipeline will have hydrate formation
conditions with free water present. The engineer then needs to specify the amount of
inhibitor needed to keep the entire pipeline in the fluid region, without hydrate
formation.
Step 1 in this calculation, the flow simulation of the pipeline, is beyond the scope of
this handbook and should be considered as a separate, pre-requisite problem, perhaps
H-0806.35
9-25
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
done by the engineering staff at the home office. As an alternative if a pipe flow
simulation is not readily available, the engineer may wish to assume that contents of a
long offshore pipeline will eventually come to the ocean bottom temperature at the
pipeline pressure.
Step 2, enabling estimations of hydrate formation pressures and temperatures, is one
of the principal goals of this handbook, as discussed in this and in the following
section. The methods below (Sections 9.3.3 and 9.3.4) may then be used directly to
determine the amount of MeOH (methanol) or MEG (monoethylene glycol) needed to
prevent hydrate formation at those conditions.
9.3.3.1
9.3.3.2
Example 4: Calculating Hydrate Formation Conditions Using the Gas Gravity Chart
Find the pressure at which a gas composed of 92.67 mol% methane, 5.29% ethane,
1.38% propane, 0.182% i- butane, 0.338% n-butane, and 0.14% pentane form hydrates
with free water at a temperature of 50 o F.
H-0806.35
9-26
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
FIGURE 9.3.3-1: HYDRATE FORMATION CURVES FOR VARIOUS GAS GRAVITY (FROM KATZ 1959)
H-0806.35
9-27
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Solution:
The gas gravity is calculated as 0.603 by the procedure below:
COMPONENT
MOL
FRACTION
yI
MOL WT
MW
AVG MOL WT
IN MIX
yI @MW
Methane
0.9267
16.043
14.867
Ethane
0.0529
30.070
1.591
Propane
0.0138
44.097
0.609
i-Butane
0.00182
58.124
0.106
n-Butane
0.00338
58.124
0.196
Pentane
0.0014
72.151
0.101
1.0000
Gas Gravity =
17.470
H-0806.35
9-28
1-Dec-00
T =
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
CW
M(100 - W)
(1)
where:
T = hydrate depression, (Teq - Toper) o F,
C =
constant for a particular inhibitor (2,335 for MeOH; 2,000 for MEG)
The Hammerschmidt equation was generated in 1934 and has been used to
determine the amount of inhibitor needed to prevent hydrate formation, as indicated in
Example 5. The equation was based upon more than 100 natural gas hydrate
measurements with inhibitor concentrations of 5 - 25 wt% in water. The accuracy of
the Hammerschmidt equation is surprisingly good; tested against 75 data points, the
average error in T was 5%.
For higher methanol concentrations ( up to 87 wt%) the temperature depression due to
methanol can be calculated by a modification of Equation (1) by Nielsen and Bucklin
(1983), where xMeOH is mole fraction methanol in aqueous phase
T = 129.6 ln(1 x MeOH )
9.3.3.4
(1a)
H-0806.35
9-29
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
100 M T
100 32 11
=
= 131
.
M T + C 32 11 + 2335
2.
3.
The amount of the free water phase is multiplied by the wt% inhibitor from the
Hammerschmidt equation, just as the inhibitor concentrations in the gas and
condensate are multiplied by the flows of the vapor and condensate. Because hydrate
inhibition occurs in the water phase, inhibitor concentrations in the gas and condensate
phases are usually counted as economic losses. Methanol recovery is done only rarely
on platforms and is typically too expensive at onshore locations.
Amount of Water Phase
The water phase has two sources: (a) produced water and (b) water condensed from
the hydrocarbon phases. The amount of produced water can only be determined by
data from the well, with an increasing amount of water production over the wells
lifetime.
Water condensed from the hydrocarbon phases may be calculated. The water content
of condensates is usually negligible, but water condensed from gases can be
H-0806.35
9-30
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
substantial. The amount of water condensed is the difference in the inlet and outlet
gas water contents, multiplied by the gas flow rate.
Rule-of-Thumb 2: For long pipelines approaching the ocean bottom temperature of
39 o F, the lowest water content of the outlet gas is given by the below table:
Pipe Pressure, psia
Water Content, lbm/MMscf
500
15.0
1000
9.0
1500
7.0
2000
5.5
An inlet gas water content analysis is used, if available. Then the water content of the
outlet gas (Rule-of-Thumb 2) may be subtracted from the inlet gas to determine the
water condensed per MMscf of gas. When an inlet gas water content is not available a
water content chart such as Figure 9.3.3-2 may be used to obtain the water content of
both the inlet and outlet gas from the pipeline.
In Figure 9.3.3-2 the temperature of the pipeline inlet or outlet is found on the x-axis
and water content is read on the y-axis at the pipeline pressure, marked on each line in
Figure 9.3.3-2. The engineer is cautioned not to use the water content chart at
temperatures significantly below 38o F. At lower temperatures the actual water content
deviates from the line due to hydrate formation. An illustration of condensed water
calculation using Figure 9.3.1-6 is given in Example 6 (Section 9.3.3.6).
H-0806.35
9-31
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
FIGURE 9.3.3-2: WATER FORMATION CURVE (FROM MC KETTA & WEHE, 1958)
H-0806.35
9-32
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-33
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-34
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-35
1-Dec-00
9.3.3.6
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
COMPONENT
MOL
FRACTION
yI
MW
AVG MOL WT
IN MIX
yI @MW
Methane
07160
16.04
11.487
Ethane
0.0194
44.09
0.855
Propane
0.0194
44.09
0.855
n-Butane
0.0079
58.12
0.459
n-Pentane
0.0079
72.15
0.570
Nitrogen
0.0596
28.01
1.670
Carbon Dioxide
0.1419
44.01
6.245
1.0000
Gas Gravity =
H-0806.35
MOL WT
22.708
9-36
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Reading the gas gravity chart Figure 9.3.3-1, the hydrate temperature is 65 o F at 1000
psia.
Step 2) Calculate the Wt% MeOH Needed in the Free Water Phase
The Hammerschmidt Equation is: T =
CW
100M - MW
100 M T
100 32 27
=
= 27
M T + C 32 27 + 2335
The weight percent of methanol needed in freewater phase is 27.0% to provide hydrate
inhibition at 1000 psia and 38 o F for this gas.
Step 3) Calculate the Mass of Liquid H2 O/MMscf of Natural Gas
Calculate Mass of Condensed H2 O
In the absence of a water analysis, use the water content chart Figure 9.3.3-2, to
calculate the water in the vapor/MMscf. The inlet gas (at 1050 psia and 195 o F) water
content is read as 600 lb m/MMscf. Rule of Thumb 2 states that exiting gas at 1000
psia and 39 o F contains 9 lb m/MMscf of water in the gas. The mass of liquid water
due to condensation is:
600 lb m _
MMscf
9 lbm
MMscf
= 591 lb m
MMscf
H-0806.35
9-37
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
day
lb H O
= 27.4 m 2
MMscf
Total Mass of Water/MMscf Gas: Sum the condensed and produced water
591 lb m + 27.4 lb m = 618.4 lb m
MMscf
MMscf
MMscf
Step 4) Calculate the Rate of Methanol Injection
Methanol will exist in three phases: water, gas, and condensate. The total mass of
methanol injected into the gas is calculated as follows:
-Calculate Mass of MeOH in the Water Phase
27.0 wt% methanol is required to inhibit the free water phase, and the mass of
water/MMscf was calculated at 618.4 lb m. The mass of MeOH in the free water phase
per MMscf is:
27wt% =
M lb m MeOH
100%
M lb m MeOH + 618.4lbm H2 O
= 228.7 lbm/MMscf
9-38
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
MeOH in Gas
= 27 lbm/MMscf
MeOH in Condensate
= 11.7 lbm/MMscf
= 267.4 lbm/MMscf
H-0806.35
1.
2.
3.
4.
9-39
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
5.
Gas Gravity - Gas gravity, calculated by the steps in Section 9.3.3.1 and Example
4.
6.
7.
8.
Formation Water Rate - Produced water flowing into the pipeline (bbl/d).
Once the above values are input, HYDCALC displays calculations for both
Intermediate Results (in black) and the amount of methanol or glycol to be injected (in
blue on a color screen). In the below example, the User Input and Calculations are
both listed in black, due to printing restrictions. A prescription for the use of this
method is shown in Example 7.
9.3.3.8
H-0806.35
9-40
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
H-0806.35
9-41
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
Solution:
Figure 9.3.3-5 on the next page is a copy of HYDCALC, highlighting the data input
that is needed to run the program. All required data are provided in the example, with
the exception of gas gravity. Gas gravity was calculated using the method described
in Example 4 to be 0.784. Figure 9.3.3-5 on the next page displays all input data and
results. The amount of methanol injected is 42.2 gal/MMscf and the amount of glycol
injected is 59.4 gal/MMscf.
For ease of use, the engineer will turn to HYDCALC to perform the second
approximation calculation. The following section provides accuracy and limitations of
both HYDCALC and the hand calculation methods, which are vital to their use.
9.3.3.9
Calculated Quantity
Water Condensed, lbm/MMscf
MeOH in Water, lb m/MMscf
MeOH in Gas, lb m/MMscf
MeOH in Condensate, lb m/MMscf
Total MeOH Injection, lbm/MMscf
Total MeOH Injection, gal/MMscf
HYDCALC
Result
591
228.7
27
11.7
267.4
40.3
619.8
239.7
24.7
11.7
276.25
42.2
While the hand calculation and the computer program provide only slightly different
results, both include inaccuracies. For example, while it is possible to obtain more
significant figures with HYDCALC than with the charts in the hand method,
HYDCALC inaccuracies are those of the charts upon which HYDCALC is based.
Using HYDCALC it was estimated that 27 wt% methanol was required in the water
phase to inhibit the pipeline, while measurements by Robinson and Ng (1986) show
that only 20 wt% methanol was required for inhibition at the same gas composition,
temperature, and pressure of Examples 6 and 7.
The major inaccuracies in the second estimation method are in the gas gravity hydrate
formation conditions, which are only accurate to 7 o F or to 500 psia. The
Hammerschmidt equation, the inhibitor temperature depression T is accurate to
H-0806.35
9-42
1-Dec-00
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
5%. With such inaccuracies, the amount of methanol or glycol injection could be in
error by 100% or more. The principal virtue of the second estimation method is ease
of calculation rather than accuracy.
A second limitation is that the method was generated for gases without H2 S, which
represents the case for many gases in the Gulf of Mexico. A modification of the gas
gravity method was proposed for sour gases by Baillie and Wichert (1987).
9.3.4
9.3.4.1
H-0806.35
9-43
1-Dec-00
9.3.4.2
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
From Windows or in the proper directory, click on, or type HYDOFF; press Enter.
2.
3.
At the Units screen, press 1 (to choose o F and psia) then Enter
4.
At the FEED.DAT question screen, press 2 and Enter if you wish to use the data in
FEED.DAT, or 1 and Enter if you wish to enter the gas composition in HYDOFF
by hand. The remainder of this example is written assuming that the user will
enter the gas composition in HYDOFF rather than use FEED.DAT. The use of
FEED.DAT is simpler and should be considered for multiple calculations with the
same gas.
5.
The next screen asks for the number of components present (excluding water).
Input 7 and Enter.
6.
The next screen requests a list of the gas components present, coded by numbers
shown on the screen. Input 1, 2, 3, 5, 7, 8, and 9 (in that order, separating the
entries by commas) and then Enter.
7.
The next series of screens request the input of the mole fractions of each
component
Methane
H-0806.35
0.7160 Enter.
9-44
1-Dec-00
Ethane
Propane
n-Butane
Nitrogen
Carbon Dioxide
n-Pentane
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
0.0473
0.0194
0.0079
0.0596
0.1419
0.0079
Enter.
Enter.
Enter.
Enter.
Enter.
Enter.
8.
9.
At the screen asking for the required Temperature, input 38, and Enter.
10. Read the hydrate formation pressure of 229.7 psia, (meaning hydrates will form at
any pressure above 230 psia at 38 o F for this gas.)
11. When asked for another calculation input 1 for No then Enter.
12. At the Options screen input 2, then Enter.
13. At the screen asking for the required temperature, input 38, and Enter.
14. At the screen to enter the WEIGHT PERCENT of Methanol, input 22.
15. Read the resulting hydrate condition of 22 wt% MeOH, 38 o F, and 1036 psia.
It may require some trial and error with the use of the program before the correct
amount of MeOH is input to inhibit the system at the temperature and pressure of the
example. One starting place for the trial and error process would be the amount of
MeOH predicted by the Hammerschmidt equation (27 wt%) in Example 6. Ng and
Robinson (1983) measured 20 wt% of methanol in the water required to inhibit
hydrates at 38 o F and 1000 psia. A comparison of the measured value with the
calculated value (22 wt%) in this example and through the Hammerschmidt equation
provides an indication of both the absolute and relative calculation accuracy.
HYDOFF can also be used to predict the uninhibited hydrate formation temperature at
1000 psia at 58.5 o F, through a similar trial and error process, as compared with 65 o F
determined by the gas gravity method. No measurements are available for the
uninhibited formation conditions of the gas in this example.
In using HYDOFF, if components heavier than n-decane (C 10 H22 ) are present, they
should be lumped with n-decane, since they are all non-hydrate formers.
H-0806.35
9-45
1-Dec-00
9.3.4.3
DEEPSTAR
MULTIPHASE DESIGN GUIDELINE
9.3.4.4