Sie sind auf Seite 1von 23

Basic Concepts of Multiphase

Flow

Prepared By

Garry A. Gregory, P.Eng.

© May 2002

Notes for a Professional Development Course


presented in

Calgary, Alberta, Canada

December 4 - 6, 2002

by

NEOTECHNOLOGY CONSULTANTS LTD.


510, 1701 Centre Street N.W.
Calgary, Alberta, Canada T2E 7Y2
Tel: (403) 277-6688 Fax: (403) 277-6687
Internet: www.neotec.com
Disclaimer

Although all of the information contained in these Short Course Notes is believed to be accurate at the time it
is prepared, it is presented without representation or warranty of any kind. Neither Neotechnology Consultants
Ltd. nor the author shall assume any liability of any kind whatsoever, either collectively or individually, arising
from the use or application of any of the technology, descriptions, or expressed opinions contained herein.
Table of Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

2. Flow Patterns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

3. Superficial Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

4. Input Volume Fractions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

5. In Situ Liquid Volume Fractions and the Liquid Holdup Effect . . . . . . . . . . . . . . . . 9

6. Actual Average Velocities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

7. Other Measures of the Liquid Holdup Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

8. Fluid Properties for Multiphase Mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

9. The Mechanical Energy Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

10. The Overall Energy Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

11. Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20


BASIC CONCEPTS OF MULTIPHASE FLOW

1. Introduction

When there is more than one fluid phase in a piping system, the complexity of any flow calculations
increases significantly. In such cases, we must take in account not only the flow rates, but also the physical
and thermodynamic properties of each phase. Interactions between the phases (e.g. phase behaviour, surface
tension) can also have a significant effect on the system behaviour. Furthermore, as will be seen in later
sections of these Notes, the elevation profile for the piping system (e.g. terrain profile for a pipeline, drilling
profile for a well) can also have a very strong effect on the pressure profile.

Throughout the rest of these Notes, it is assumed that we always have adequate models for all
required PVT behaviour and fluid transport properties. In other words, for any particular combination of
pressure and temperature, we either know or can readily compute the flow rate, density, viscosity, specific
enthalpy etc. for any phase that exists. Procedures for doing this have been discussed in some detail in
Chapter 2.

A detailed discussion and derivation of the rigorous mass, momentum, and energy balance equations
that can be written is beyond the scope of these Notes. In any case, as far as oil and gas industry applications
are concerned, the main value of these equations is limited, for all but a few specialized cases, to the insight
they provide with respect to the importance of various terms and constraints that must be satisfied. For the
most part, it is either impossible or impractical to solve the equations rigorously because of the number of
unknowns, the complexity of the boundary conditions, parameter interactions, and the fluid systems
themselves.

Historically, this has meant that we were limited in large measure to performing design and analysis
studies using purely empirical correlations. While most of these correlations were originally based on small
scale laboratory data, many have been demonstrated to be considerably more robust than would otherwise
be expected, based on extensive testing with actual field data from a wide variety of systems.

Between the extreme complexity of the rigorous theoretical equations and the relative simplicity of
the empirical correlations lies what we refer to as Mechanistic models. These models are usually developed
to describe the behaviour of the system over a limited range of conditions for which some particular fluid
transport mechanism is assumed to prevail. A set of equations, postulated to describe that mechanism, is
written and solved according to appropriate assumptions. Such models usually involve a mixture of rigorous
development, classical fluid mechanics (either theoretical or empirical), and purely empirical relations. They
can be quite complex to use, but can also be capable of very good accuracy and can sometimes even predict
the characteristics of a given flow in some detail. While proven mechanistic models are not available for all
cases of interest, their development and use is steadily increasing. Because of their potential for greater
accuracy, it is reasonable to expect that they will replace most of the purely empirical methods in the not
distant future. We will look at a number of these in some detail in later Chapters of these Notes.

Regardless of the model or procedure used, there are a number of specialized concepts and parameter
definitions for multiphase flow that apply to both pipelines and wells. The remainder of this Chapter deals
with those that will be encountered throughout the rest of these Notes.
Basic Concepts of Multiphase Flow (Revised 05-02) Page 2

2. Flow Patterns

One of the most obvious characteristics of multiphase flow is the occurrence of different flow
patterns (sometimes referred to as flow regimes). Depending on the system, the flowing mixture can range
from a continuous gas phase with a minute quantity of liquid to a continuous liquid phase with very small
amounts of gas. Between these limits of essentially single phase flow, a wide variety of flow patterns (i.e.
spatial distributions of gas and liquid in the flowing mixture) can occur.

Unfortunately, there is no standardized and unambiguous way to classify the different flow patterns.
Different authors have used a wide variety of descriptive terms to describe their visual observations, and
around one hundred such terms have reportedly appeared in the literature. For example, one particular flow
pattern may be described variously as elongated bubble, plug, intermittent, slug I, semi-slug, and so on.

In these Notes, we will arbitrarily define six flow patterns for horizontal flow and four for vertical
flow, as shown schematically in Figures 2.1 and 2.2 respectively. These represent the primary flow patterns,
from a mechanistic sense, that can exist under stable, steady-state conditions. Most of the additional terms
that have been used in the literature represent attempts by authors to characterize transition regions between
the primary flow regimes.

Figure 2.1

Basic Flow Patterns for Multiphase Flow in Horizontal Pipes


Basic Concepts of Multiphase Flow (Revised 05-02) Page 3

In horizontal flow, for example, the Stratified, Wave, and Elongated Bubble flow patterns shown in
Figure 2.1 are primarily gravity dominated, although interfacial shear has a significant effect in Wave flow.
By contrast, shear and momentum effects tend to dominate in Slug, Annular-Mist, and Dispersed Bubble
flow. As might be expected, the flow patterns in the first group are much more affected by even small
changes in the inclination angle of the pipe than are those in the latter group. This has a significant impact
on flow pattern prediction in cross country pipelines, which are never completely horizontal.

Figure 2.2

Basic Flow Patterns for Multiphase Flow in Vertical


Pipes

In vertical flow, gravity effects are essentially uniform over the cross section of the pipe, and, as can
be seen in Figure 2.2, flow patterns thus tend to be axially symmetrical. Drag forces on bubbles and droplets
tend to be significant, and turbulence plays a major role in determining maximum stable bubble and/or
droplet sizes.

As we shall see later, transitions from one flow pattern to another are strongly influenced by changes
in the gas and liquid flow rates. Such changes might occur because of additional fluids being introduced into
the system, but, as we saw in Chapter 2, the volumetric flow rate of a given phase can change significantly
Basic Concepts of Multiphase Flow (Revised 05-02) Page 4

in response to pressure and temperature changes, even though there is no net change in the total mass flow.
It is thus common for the flow pattern to change, sometimes quite dramatically, as the fluid moves along the
pipe. Since pressure changes can be large as a result of elevation changes, this is an especially important
factor that must be considered when performing calculations for wells and production risers.

The wide variation in the possible flow patterns that can occur might intuitively be expected to
reflect the different energy dissipation mechanisms that can take place and therefore, be related to pressure
losses. It is thus important to be able to predict what flow pattern will occur in a given case, and a number
of methods for doing this have been reported in the literature. Several of these are discussed in later Chapters
of these Notes

3. Superficial Velocity

The superficial velocity of a given phase is the velocity based on the volumetric flow rate of the
phase and the total cross-section area of the pipe. Thus, for the liquid phase,

(3-1)

where VSL = superficial liquid velocity (ft/s or m/s)


A = cross sectional area of the inside of the pipe (ft2 or m2 )
QL = volumetric liquid flow rate at flowing P, T (ft3 /s or m3 /s)
D = pipe diameter (ft or m)

Similarly, for the gas phase,

(3-2)

where VSG = superficial gas velocity (ft/s or m/s)


QG = volumetric gas flow rate at flow P, T (ft3 /s or m3 /s)

Note that since QL and QG are volumetric flow rates at the flowing pressure and temperature, V SL and V SG
are also functions of pressure and temperature and will thus change throughout the system.

A related parameter is the average mixture velocity, defined as

(3-3)

where VM = average mixture velocity (ft/s or m/s)

Note that while VSL and VSG are not, in general, the actual average velocities of the liquid and gas phases
respectively, VM is the actual average mixture velocity, since it is based on the total volumetric flow.

It should also be noted that, since we assumed that we have satisfactory fluid PVT behaviour models,
Basic Concepts of Multiphase Flow (Revised 05-02) Page 5

it can also be assumed that we always know QL and QG . Clearly, this means that we can always calculate VSL ,
VSG , and VM for any pressure and temperature, without knowing anything else about the flowing system.

4. Input Volume Fraction

The input volume fraction of a given phase is defined in terms of the volumetric flow rates of each
phase at the flowing pressure and temperature. Thus, for the liquid phase, we have

(4-1)

and for the gas phase,

(4-2)

If Equations (3-1) and (3-2) are substituted into (4-1), we can obtain the relation,

(4-3)

and if we further substitute Equation (3-3) into Equation (4-3), we obtain the alternate relation,

(4-4)

Similarly, it is easily demonstrated that

(4-5)

In Section 3, it was noted that we can always compute VS G , VS L and VM for any pressure and
temperature. It is obvious from Equations (4-1) through (4-5) that this is also true for CL and CG .

5. In Situ Liquid Fraction and the Liquid Holdup Effect

In general, the fraction of the pipe that is occupied by the liquid phase under steady state flowing
Basic Concepts of Multiphase Flow (Revised 05-02) Page 6

conditions is not the same as the input liquid fraction. To illustrate why this is so, we will consider a simple
example involving cars on a highway.

For simplicity, it is assumed that there are only two kinds of vehicles that will use the highway, as
shown in Figure 5.1. These are:

Figure 5.1

Vehicle Types for the Highway Analysis Problem

Figure 5.2

Feeder Section and Main Section for Highway Analysis Problem

(a) Type "L" vehicles, which always travel at exactly 60 km/hr.


(b) Type "G" vehicles, which always travel at exactly 120 km/hr.

The vehicles are assumed to enter the highway system through a single feeder lane section, as shown
in Figure 5.2, such that six Type "G" vehicles enter for every four Type "L" vehicles. The input fractions
can thus be defined as,
Basic Concepts of Multiphase Flow (Revised 05-02) Page 7

We can also define QL as the total number of Type "L" vehicles that enter the system per hour, along with
a corresponding definition for QG . Then, it is also clear that,

Now, the total time taken by a type "L" vehicles to pass through a particular stretch of highway is
simply the distance divided by the velocity, and thus, for some length )L,

Similarly,

At any given instant, the number of type "L" vehicles that will be on the highway in the length )L (i.e. the
holdup ) is simply the product of the number of vehicles per unit time and the transit time, or, in this case,

The total number of both types of vehicles in a distance of )L along the highway will therefore be,

The in situ fraction (i.e. the actual fraction on the highway) of type "G" vehicles, say, EG , is the number of
type “G” vehicles, divided by the total number of vehicles, or,
Basic Concepts of Multiphase Flow (Revised 05-02) Page 8

We know, however, from the input specifications of the system that,

and thus,

Similarly,

Recall, however, that CG = 0.6 and CL = 0.4 . It is thus apparent that:

(i) for the slower vehicle, the in situ fraction is greater than the input fraction
(ii) for the faster vehicle, the in situ fraction is less than the input fraction

In a multiphase pipeline, the density and viscosity of the gas phase are generally both much lower
than for the liquid phase(s), and the mobility of the gas is therefore much greater. For a given pressure
gradient, the velocity of the gas will thus be higher than that of the liquid phase(s). The exceptions to this
occur for flow patterns where the flow is pseudo-homogeneous (Annular-Mist, Dispersed Bubble). For
pseudo-homogeneous flows, the in situ (or actual average) velocities of both phases are essentially the same.
It is easily demonstrated using the above analysis that in such cases EL = CL (and thus, EG = CG ).

In most cases however, the gas velocity is greater than that of the liquid, and the in situ liquid
fraction will thus be greater than the input liquid fraction. This phenomenon is often described as the holdup
effect or the slip effect (i.e. the gas phase slips past the liquid phase).

The flow pattern and pressure gradient can logically be expected to depend on the conditions which
exist in the pipe, and it is thus reasonable to assume that the in situ liquid fraction is an important parameter
Basic Concepts of Multiphase Flow (Revised 05-02) Page 9

of interest. In fact there are several possible definitions for the in situ volume fraction and we will discuss
them briefly below:

Local average in situ liquid volume fraction (,L )

This is the time-averaged fraction of liquid in a small volume increment (theoretically, at a single
point) at some fixed location in the pipe, as depicted in Figure 5.3 It is typically measured using a small
probe (e.g. conductivity, optical). This is an important parameter in heat transfer applications (especially
in the nuclear industry) since a local "dry spot" can lead to very high heat flux and even melting of the tube
wall. For petroleum industry applications the local average in situ liquid fraction has historically been of
little interest. That, however, is changing to some extent because of studies that have been carried out to
determine why some sections of a pipe exhibit extremely rapid corrosion rates. Evidence suggests that in
some flow regimes, most notably in Slug flow, the local liquid fraction can be an important factor in being
able to predict potential problem sites. It is reasonable to assume that more attention will be paid to this
parameter as a consequence, since pipeline failures due to corrosion represent a growing problem for the oil
and gas industry.

Figure 5.3

Local Average In Situ Liquid Volume Fraction, , L

Cross section average in situ liquid volume fraction (EL )

The cross-sectional average in situ liquid volume fraction as depicted in Figure 5.4, can be defined
as,

(5-1)

where A L = average cross-section area of the pipe occupied by the liquid phase
,L = local average in situ liquid volume fraction

The cross section average liquid volume fraction can be measured using a local probe that has a traversing
mechanism (i.e. physically measure ,L at various points and perform the integration numerically). It is more
common however to use a device that measures the cross section average liquid fraction directly (e.g. gamma
or neutron densitometer, x-ray). As we shall see shortly, this parameter is of particular interest in oil and gas
Basic Concepts of Multiphase Flow (Revised 05-02) Page 10

industry applications.

Figure 5.4

Cross Section Average


In Situ Liquid Volume
Fraction

Volume average in situ liquid volume fraction (L )

The volume average in situ liquid volume fraction as depicted in Figure 5.5, can be defined as,

(5-2)

In laboratory studies, L was historically measured using two quick closing valves to trap the liquid in a given
section of pipe. A third valve allowed the liquid to be drained, collected, and measured.

Figure 5.5

Volume Average In Situ Liquid Volume Fraction

For actual pipelines,  L can be very difficult to measure with any accuracy. It is sometimes possible
to measure EL at a number of selected locations, and then perform a numerical integration. Alternatively,
it may be possible to send a pigging sphere through the line and measure the liquid that is either collected
in a downstream separator, or metered over a period of time.
Basic Concepts of Multiphase Flow (Revised 05-02) Page 11

The distinction between EL and  L is thus that while E L is defined for the cross section at a single
point along the pipe,  L is defined as the average over some fixed length of pipe. Recall however that the
Stepwise calculation procedure is essentially a numerical integration analysis. Thus, if a relatively short
calculation segment is selected, there is no practical difference between EL and  L . For this reason, we will
make no further distinction in these Notes between the two definitions; the context should make clear the
meaning intended in a given instance.

As was demonstrated earlier, the in situ fractions differ from the input fractions when the velocities
of the phases are not the same. In general, we do not know what the actual velocity of a given phase will be,
relative to the other phase(s), and must depend on an appropriate correlation or model to estimate that. Thus,
while we can always calculate the input volume fraction directly from the fluid behaviour models, the in situ
fraction depends directly on the multiphase flow technology that is selected.

6. Actual Average Velocities

While the superficial velocity is based on the entire cross section area of the pipe, the true or actual
average velocity of a phase is based on the volumetric flow rate of the phase at the flowing pressure and
temperature and the actual cross section area of the pipe occupied by that phase. Thus, for the liquid phase,
we can write,

(6-1)

If Equations (5-1) and (3-1) are substituted into Equation (6-1), we obtain the relations,

(6-2)

Similarly,

(6-3)

The true average velocities must be known to compute the liquid and gas transit times through the
pipe, and they are also used when evaluating possible erosion effects. Since they are directly related to the
in situ volume fractions, they also depend on the particular multiphase flow model that is selected.

7. Other Measures of the Liquid Holdup Effect

The in situ liquid volume fraction is the primary parameter that we need to know when designing
a new facility, or analysing the performance of an existing facility. However, it is not the only parameter that
can be used to characterize the liquid holdup effect. Several others that will be found in the literature are
discussed briefly below:
Basic Concepts of Multiphase Flow (Revised 05-02) Page 12

Slip Velocity (S)

The slip velocity is defined as the difference between the true average velocities, i.e.,

(7-1)

The substitution of Equations (6-2) and (6-3) into (7-1) results in,

(7-2)

which, after some algebraic manipulations, can be rearranged to give,

(7-3)

Note that if the average slip velocity is zero, there is no holdup of one phase relative to the other, i.e.

which can be solved for EL to give,

Thus, when the slip velocity is zero (i.e. both phases are moving at the same actual average velocity), the in
situ fractions are equal to the input fractions and are thus determined solely by input parameters. This is true
(or at least approximately so) for pseudo-homogeneous flow regimes (i.e. Dispersed Bubble, Annular-Mist).
Because of this, multiphase flow models that assume pseudo-homogeneous flow, or that base the calculation
of mixture properties on the input volume fraction, are often referred to as no-slip models.

Holdup Ratio (H)

The holdup ratio is defined as the ratio of the true average velocities, i.e.

(7-4)

It is easily verified that the following expressions are also equivalent,


Basic Concepts of Multiphase Flow (Revised 05-02) Page 13

(7-5)

One can also solve Equations (7-1) and (7-4) to obtain, after some algebra,

(7-6)

Bankoff K Factor

The Bankoff K Factor is defined as the ratio of the gas in situ volume fraction to the gas input volume
fraction, i.e.,

(7-7)

While most of the papers in the literature contain correlations and models that have been expressed
in terms of EL (or EG ), all of the other parameters (S, H, and K) discussed above have also been used on
occasion. Clearly, however, they are all inter-related and all relate to the same multiphase flow phenomenon.
The ability to predict any one of them in terms of known parameters is sufficient to determine any of the
other three, as may be required.

8. Fluid Properties for Multiphase Mixtures

In most multiphase flow calculation procedures, it is necessary to assign values for various transport
properties to the flowing mixture, such that, in some way, the contributions of all phases are acknowledged.

Before proceeding further however, it must be noted that, with few exceptions, multiphase flow
models are, in reality, two phase flow models. That is, they simply assume that the system consists of a gas
phase and one liquid phase. In many cases, however, there are actually two liquid phases; a hydrocarbon
liquid phase and a predominantly water phase. At this time, however, the only way to account for two liquid
phases is these procedures is to assume that the two liquids exist as a pseudo-homogeneous mixture and can
be dealt with as a single liquid phase having properties that somehow reflect contributions of both liquids.
The most common ways of defining the properties of this single pseudo-homogeneous liquid are given by
the expressions below.

First of all, since the phases are immiscible, the volumetric flow rates must be additive, and the
effective liquid flow rate is thus given by,

(8-1)

where QL = volumetric flow rate of the "liquid phase" at flowing conditions (ft3 /s or m3 /s)
QO = volumetric flow rate of oil or hydrocarbon liquid phase at flowing conditions (ft3 /s
or m3 /s)
QW = volumetric flow rate of water phase at flowing conditions (ft3 /s or m3 /s)
Basic Concepts of Multiphase Flow (Revised 05-02) Page 14

Since the liquid phase is assumed to be a pseudo-homogeneous mixture, its properties will be based
on the input water fraction (in the liquid phase only), or

(8-2)

The effective density of the liquid phase is computed as a weighted average of the hydrocarbon liquid
and water phase densities, or,

(8-3)

where DL = effective liquid phase density at flowing conditions (lb/ft3 or kg/m3 )


DW = water phase density at flowing conditions (lb/ft3 or kg/m3 )
DO = hydrocarbon liquid phase density at flowing conditions (lb/ft3 or kg/m3 )

The effective viscosity of the liquid phase, assuming that no emulsion occurs, is usually computed
using either the relation,

(8-4)

or

(8-5)

where :L = effective viscosity of the liquid phase at flowing conditions (cP or mPa.s)
:w = viscosity of the water phase at flowing conditions (cP or mPa.s)
:O = viscosity of the hydrocarbon liquid phase at flowing conditions (cP or mPa.s)

Finally, the effective specific enthalpy of the liquid phase is given by,

(8-6)

where HL = effective specific enthalpy of the liquid phase at flowing conditions (BTU/lb or
kJ/kg)
HW = specific enthalpy of the water phase at flowing conditions (BTU/lb or kJ/kg)
HO = specific enthalpy of the hydrocarbon liquid at flowing conditions (BTU/lb or kJ/kg)

If no water phase is present, all of the liquid phase properties are identical to those of the
hydrocarbon liquid, since CW = 0 . Unless stated otherwise, however, it will always be assumed that the
presence of a water phase is implicitly taken into account, as illustrated above, in the flow rate and properties
ascribed to the liquid phase.
Basic Concepts of Multiphase Flow (Revised 05-02) Page 15

We can now proceed with the definitions for the properties of our multiphase (gas-liquid) mixture.

Density

The mixture density is defined in terms of the in situ volume fractions and the densities of the
individual phases, as follows:

(8-7)

where DM = effective mixture density at flowing conditions (lb/ft3 or kg/m3 )


DL = effective density of the liquid phase at flowing conditions (lb/ft3 or kg/m3 )
DG = density of the gas phase at flowing conditions (lb/ft3 or kg/m3 )

This is a true measure of the average in situ density and, as we shall see shortly, it is an especially important
parameter in the calculation of overall pressure losses.

Many analyses are also based on a homogeneous flow or no-slip model, as noted earlier. We thus
have another similar definition for the density of the mixture, based on the input volume fraction i.e.,

(8-8)

where DNS = no-slip mixture density at flowing conditions (lb/ft3 or kg/m3 )

Viscosity

While D M , as defined by Equation (8-7) is a valid expression for the actual average density of the
mixture (as is D NS when the no-slip model assumptions apply), no simple expression exists to rigorously
define the actual viscous behaviour of a gas-liquid mixture. Nevertheless, in many instances, mixture
viscosities calculated using expressions analogous to Equations (8-7) and (8-8) have been found to give
useful results. Thus, we have

(8-9)

and
Basic Concepts of Multiphase Flow (Revised 05-02) Page 16

(8-10)

where :M = effective mixture viscosity at flowing conditions (cP or mPa.s)


:NS = "no-slip" mixture viscosity at flowing conditions (cP or mPa.s)
:L = effective viscosity of the liquid phase at flowing conditions (cP or mPa.s)
:G = viscosity of the gas phase at flowing conditions (cP or mPa.s)

As for oil-water mixtures, there are two other expressions that are often encountered, i.e.

(8-11)

(8-12)

To re-iterate, both :M and :NS , as calculated using any of the above equations, must be viewed as
hypothetical transport properties that are useful for modelling purposes, rather than as representative of the
actual viscosity of the mixture.

Specific Enthalpy

In any heat transfer calculation, the change in enthalpy is important; the absolute enthalpy in itself
is of little interest. Since we will always be dealing with a steady state flow process, the only definition of
mixture enthalpy that we require is based on the input volume fraction. Thus,

(8-13)

where HM = effective specific enthalpy of the mixture at flowing conditions (BTU/lb or kJ/kg)
HL = effective specific enthalpy of the liquid phase at flowing conditions (BTU/lb or
kJ/kg)
HG = specific enthalpy of the gas phase at flowing conditions (BTU/lb or kJ/kg)

9. The Mechanical Energy Equation

As noted earlier in Chapter 4, a detailed derivation of the mechanical energy equation is beyond the
scope of these Notes, and the result will simply be stated here. For some finite length of )L, inclined at an
angle 2 from the horizontal, and assuming that all fluid properties are constant over the length of )L (or
nearly so), the mechanical energy equation for a two phase (or pseudo two phase) mixture is,
Basic Concepts of Multiphase Flow (Revised 05-02) Page 17

(9-1)

where )P = total pressure drop over the length of )L (psi or kPa)


)L = length of pipe section (ft or m)
WG = mass flow rate of the gas phase (lb/s or kg/s)
WL = mass flow rate of the liquid phase (lb/s or kg/s)
VG1 = average in situ velocity of the gas phase at the upstream end of the pipe section (ft/s
or m/s)
VL1 = average in situ velocity of the liquid phase at the upstream end of the pipe section
(ft/s or m/s)
VG2 = average in situ velocity of the gas phase at the downstream end of the pipe section
(ft/s or m/s)
VL2 = average in situ velocity of the liquid phase at the downstream end of the pipe
section (ft/s or m/s)
"G = gas phase velocity profile correction factor (typically about 0.8 to 0.9 for turbulent
flow)
"L = liquid phase velocity profile correction factor (typically about 0.8 to 0.9 for
turbulent flow)
g = gravity constant (32.2 ft/s2 or 9.81 m/s2 )
gc = dimensional conversion factor (32.2 ft lbm /lbf-s2 or 1.0)
2 = angle of inclination relative to the horizontal (degrees)

In spite of its apparent complexity, just as in the case of single phase flow, each term on the right hand side
of Equation (9-1) represents a particular mode of energy change, and, for simplicity, the equation can be re-
written as,

(9-2)

For single phase flow, )PPE is the pressure loss due to potential energy effects, and it is also
legitimately a hydrostatic head effect (i.e. it can be computed simply on the basis of the average density of
the fluid). To be consistent with the fundamental laws of thermodynamics, )PPE must represent a reversible
energy change. Flow uphill results in an increase in potential energy at the expense of a decrease in pressure.
Downhill flow back to the original datum must therefore result in an exactly equal decrease in potential
energy.

It is, however, a well documented fact for multiphase systems that the accompanying increase in
pressure for downhill flow is typically less than the pressure decrease during a corresponding uphill flow
because of irreversible losses, especially at very low flow rates. Unfortunately, those irreversible losses are,
at least in part, due to the fact that the flow pattern (and thus, the liquid holdup) is generally different in
uphill flow than in downhill flow. A consequence of this, if we follow the usual logic and assume that )Pf
is the only irreversible term, is that our model or correlation for )Pf must be extremely complex, since it has
to account for all elevation and flow regime effects, as well as what we customarily refer to as "friction" (i.e.
velocity related) losses. We thus have two terms that are effected by elevation changes, and our ability to
Basic Concepts of Multiphase Flow (Revised 05-02) Page 18

examine the importance to the overall pressure loss of individual effects is severely reduced.

Fortunately, it turns out that we can eliminate this problem by re-writing Equation (9-2) in terms of
a hydrostatic component, rather than a potential energy component. i.e.,

(9-3)

where )PHH = pressure loss due to hydrostatic head effects (psi or kPa)
)Pf* = pressure loss due to friction effects (psi or kPa)

As a recommended procedure, the hydrostatic head loss in uphill flow is calculated using an
expression that is similar to the standard equations for computing potential energy change, but it is based on
the overall in situ mixture density , rather than on the mass throughput in the system, i.e.,

(9-4)

For downhill flow, if we use the mixture density, the pressure recovery tends to be over-estimated.
Much better results are generally obtained if we use the expression,

(9-5)

i.e the pressure recovery is based on the density of the gas phase.

For horizontal flow, the inclination angle is zero, and,

(9-6)

It must be noted, however, that Equation (9-3) in effect defines a pressure change due to irreversible
effects (i.e. )Pf* ) that excludes the effect of liquid holdup on mixture density, and which must therefore
differ from the )Pf term that is defined by Equations (9-1) and (9-2). In other words, all of the elevation
related pressure loss effects, including the elevation effect irreversibilities, have been transferred to )PHH .
The relationship between the two "friction" loss terms is readily shown to be,

(9-7)

Equation (9-3) is, in fact, the basis of all of the multiphase flow methods and models that are
Basic Concepts of Multiphase Flow (Revised 05-02) Page 19

discussed in subsequent chapters of these Notes. For simplicity, we will drop the superscript from the )Pf*
term throughout the remainder of these Notes, and will just write the equation as,

(9-8)

The reader should keep in mind however the distinction between )PPE and )PHH . The former is the
totally reversible pressure change due to elevation effects, while the latter is generally treated as an only
partially reversible term.

One final comment must be made regarding )PHH . In earlier days of multiphase flow computations,
it was common practice to compute )PHH using Equations (9-4) for uphill flow. However, for downhill flow,
it was simply assumed that )PHH = 0. This very often lead to an over-estimate of the pressure loss,
especially in gas-condensate or other low liquid systems. The practice we currently recommend, for all but
a few specialized cases, is to compute )PHH using Equations (9-4) through (9-6), as outlined above.

For downhill flow, Equation (9-8) will thus predict a pressure recovery based on the density of the
gas phase only. At low pressures, or high flow rates (where the friction losses are high), the magnitude of
this pressure recovery will generally be small relative to the overall pressure loss, and therefore, not greatly
different from ignoring it altogether. However, for gas-condensate pipelines, especially at higher pressures,
the pressure recovery effect can be very significant, and taking it into account in this way will generally
improve the accuracy of the pressure loss calculations considerably.

10. The Overall Energy Equation

For two phase flow, the overall energy balance over a pipe section having the length )L can be
written as,

(10-1)

and
(10-2)

(10-3)

where q = rate of heat transfer from the flowing fluid to the pipe surroundings (BTU/s or W)
Uo = fluid-to-surroundings overall heat transfer coefficient based on the outside diameter
of the pipe (BTU/ft2 -s-o F or W/m2 -K)
Do = outside diameter of the pipe (ft or m)
)L = length of pipe section (ft or m)
Tf = average temperature of the flowing fluids in the pipe section (o F or o C)
Ts = average temperature of the surroundings for the pipe section (o F or o C)
GM = mass flux of the flowing two phase mixture (lb/ft2 -s or kg/m2 -s)
2 = pipe angle of inclination relative to the horizontal (degrees)
" = kinetic energy velocity profile factor (typically 0.8 to 0.9 for turbulent flow)
J = energy units conversion factor (777.6 BTU/ft-lbf or 9.81 m/s2 )
Basic Concepts of Multiphase Flow (Revised 05-02) Page 20

The subscripts 1 and 2 on the mixture enthalpy HM and mixture density D M indicate that these
parameters are to be evaluated for conditions at the upstream and downstream ends of the pipe section
respectively. All other parameters are assumed to be evaluated at the average conditions in the pipe section,
in the usual manner when the Stepwise calculation procedure is used.

11. Concluding Remarks

Just as for single phase flow, the calculation of both the pressure and temperature profile along the
pipe involves the simultaneous solution, using the Stepwise procedure, of the mechanical energy and overall
energy equations. This, of course, can be a complex iterative procedure in each incremental section of pipe
for a two phase system. In general terms, the requirements are as follows:

(i) The phase behaviour and all fluid properties must be re-evaluated at each iteration to adequately
account for pressure and temperature effects.

(ii) The in situ liquid volume fraction (used to compute )PHH ) and the friction loss term )Pf must be
re-evaluated at each iteration using an appropriate multiphase flow model.

(iii) The length of pipe section which constitutes each step (i.e. )L) must be selected such that it is
reasonable to assume that all fluid properties are approximately constant within the section.

All of the concepts, definitions, and equations presented in this Chapter are completely general and
apply equally to pipelines and wells. Obviously, the individual terms in equations can be expected to have
greater or lesser importance as a function of the geometry of the pipe. For example, in a horizontal (or nearly
so) pipeline, )PHH will be negligible. For pipeline systems in very hilly terrain, )PHH can easily be the same
order of magnitude as )Pf , especially at lower overall flow rates when liquids tend to accumulate at low
points. In an oil well, )PHH typically accounts for 95% or more of the total pressure drop.

In most cases, )PKE will be negligible with respect to the overall pressure loss (typically less than
1%). The exceptions occur when overall pressure loss is large relative to the absolute pressure (i.e. > 50%).
This almost invariably happens only at relatively low pressures, during a blowout or blowdown situation,
or with extremely high flow rates..

In any case, careful examination of the relative magnitude of the terms on the right hand side of
Equation (9-6) can often be useful when analysing the expected performance of an existing or proposed
system under some proposed set of conditions.

In subsequent Sections of these Notes, we will look at individual models and correlations for
performing the multiphase flow calculations, with emphasis on when to use a given procedure.

Das könnte Ihnen auch gefallen