You are on page 1of 30

Chapter 2 Definitions and basic quantities - Version 2017/1 11

Chapter 2
Definitions and basic quantities

2.1 Single phase fluid properties


In petroleum production the phase behaviour of gas and oil is essential for the flow
characteristics in pipes. This course primarily covers the description of flow on the premise
that the fluid properties are already known. Nevertheless we will have an introductory look at
interfacial tension before we turn to the description of two-phase fluid behaviour.

2.1.1 Interfacial tension


Interfacial tension σ is energy pr. unit surface or interface area. We normally discriminate
between interfacial tension between two phases, and the surface tension on the surface of a
medium against vacuum. Often surface tension is used for the liquid and its vapour phase.

The units are: [σ] = [Energy pr. unit area] = Jm-2 =Nm-1 = [Force pr. unit length]

Energy pr. unit area is often used when the interfacial tension is considered as a
thermodynamic quantity, while force pr. unit length is used when we consider it as a tension
or a force. The surface tension for some fluids is given in the table below. For liquid-vapour
systems at critical point, the surface tension approaches zero. Contamination influences the
interfacial tension. For fluids which are not completely isolated, one observes that even short
time after the generation of an interface, the interfacial tension is reduced, due to
accumulation of foreign substances which may accumulate at the surface. Even electric
charge may change the interfacial tension. This can be seen e.g. in plastic pipes carrying oil
and air. At low flow velocities one may observe “tears” and “fingers” along the wall which
are due to Marangoni convection.

Fluid system Interfacial tension [mN/m]


Water-air 72.2 (at T= 15 0C, P=1atm)
Mercury-air 300
High-pressure oil-gas Often <10

Table 2.1 Surface tension for some fluid-fluid systems

The physical basis of interfacial tension is associated with different “population density” of
each molecule type in each phase. Also there is and increasing intermolecular distance as one
moves from the liquid phase through the interface to the gas phase. The Lennard-Jones
potential U(r) varies with the distance r, and the force
 ∂U (r ) 
F (r ) = − ir
∂r
becomes attractive when the molecules are separated beyond the minimum of the potential.

In the following is discussed some consequences of the interfacial tension.


12 Chapter 2 Definitions and basic quantities - Version 2017/1

U(r)
Attraction domain

Repulsion
domain
dU
dr

Intermolecular distance r

Molecule 1 Molecule 2

Figure 2.1 The Lennard-Jones intermolecular potential from molecule 1 on molecule2.

Laplace’s equation
Laplace’s equation states that
 1 1 
pi = p o + σ  + 
 R1 R2 
where pi and po are the pressures at the “inside” and the outside of a curved interface. As
shown in figure 2.2, R1 and R2 are the principal radii on an equivalent ellipsoid made to match
the interface.

For a spherical interface R2 = R1 = R, thus pi = p o +
R

Capillary forces
It is well known that when a capillary tube is immersed into a liquid, the liquid level inside
the tube is higher than outside. This is a consequence of Laplace’s equation. We can calculate
the height difference H in a tube with radius R by assuming that the gas-liquid interface inside
the pipe is semi-spherical. At equilibrium we have

hydrostatic pressure difference 


pressure difference across interface
   2σ 2σ
( ρ L − ρ G ) gH = ⇒H =
R ( ρ L − ρ G ) gR

A more accurate analysis shows that the contact angle ψ between liquid and pipe wall is
2σ cosψ
important, and the corresponding equation is: H =
( ρ L − ρ G ) gR
Chapter 2 Definitions and basic quantities - Version 2017/1 13

R1

R2

Figure 2.2 The principal radii on an ellipsoidal bubble

Youngs equation
Youngs equation are used for equilibrium calculation of complex interface, e.g. an oil drop
floating on a water surface as shown in figure 2.4. The interfacial tension force can be
imagined to act tangentially to the interface at any position. At the intersection between all the
three interfaces (which is a ring around the oil drop in the horizontal plane) the forces on each
interface must balance at equilibrium. Decomposing the forces in the vertical and horizontal
directions we then get:

Horizontally: σ WG = σ WO cosθ 1 + σ OG cosθ 2


Vertically: σ WO sin θ 1 = σ OG sin θ 2

From these two equations we may determine the two angles provided the interfacial tensions
are known.

Exercise: Show that the angles θ1 and θ2 are given by cos(θ ) =


(σ ) + (σ ) − (σ )
WG 2 WO 2 OG 2

2σ WG ⋅ σ WO
1

and cos(θ ) =
(σ ) + (σ ) − (σ )
WG 2 OG 2 WO 2

2σ WG ⋅ σ OG
2

2.1.2 Physical effects of interfacial tension in two phase flow

Capillary forces influence the speed of surface waves

Surface waves on a liquid having density ρ have a speed given by


gλ 2πσ
v= +
2π ρλ
where λ is the wavelength and g is acceleration of gravity. The wave velocity is dependent on
the “water depth”, and this particular expression refers to “deep water” waves.
14 Chapter 2 Definitions and basic quantities - Version 2017/1

ρG H

ρL

Figure 2.3: Liquid level rise in capillary tubes

GAS
σ OG
σ WG θ2 OIL
θ1
WATER
σ WO

Figure 2.4: Youngs equation – force equilibrium


Chapter 2 Definitions and basic quantities - Version 2017/1 15

Capillary

Gas bubbles in capillaries


may become stuck

ρL σ g ∆ρ 1 / 4
U 0 = 1.53 [ ]
ρL2
ρG

Figure 2.5: The rise velocity of bubbles in liquid depends on surface tension. Large bubbles
easily become irregular and are more influenced.

The rise velocity of bubbles is modified by interfacial tension.

Medium sized bubbles (d ≈ 1mm – 1cm) have a rise velocity given by

σ g ( ρ L − ρ G ) 1/ 4
U 0 = 1.53 [ ]
ρL2

The interfacial tension is important for the shape of the bubbles. In thin capillaries the bubbles
may stop up completely

Exercise: Explain the influence of interfacial tension on the rise velocity. What happens for
medium sized bubbles if the interfacial tension becomes very small. Why are gas bubbles
stuck in small capillaries?

Bubble size and drop size in turbulent flow depend on interfacial tension

We will see in chapter 4 that the largest stable bubble in turbulent can be modelled in a simple
way by the non-dimensional Weber number defined by
ρU 2 d
We =
σ
which expresses the relation between dynamic (turbulent) forces and surface forces.
16 Chapter 2 Definitions and basic quantities - Version 2017/1

He
1

H2
N2
0.9
σ (under pressure)/ σ (pure water)

0.8

C2H6
0.7

0.6

CO2
0.5
0 20 40 60 80 100

Pressure, atm

Figure 2.6: Effect of gas pressure on surface tension of a water surface at 25 0C.
[E.J. Slowinski, Jr., E.F. Gates, and C.E. Waring, J. Phys.Chem., 61:808 (1957)]

2.1.3 The dependence of interfacial tension on pressure and temperature


Interfacial tension is more sensitive to temperature than to pressure. A common correlation to
relate interfacial tension between two reduced temperatures Tr1 and Tr2 is given by

1.2
σ1  1 − Tr1 
  (T < Tc )
σ2 1 − Tr 2 
T
where the reduced temperature is given by Tr = and Tc is the critical temperature. Note
Tc
that for liquid the reduced temperature is less than 1. The interfacial tension decreases with
increasing temperature of the liquid. The interfacial tension varies with pressure by two
mechanisms, that both contribute to reducing σ when pressure increases:
1. Pressure increase increases the amount of dissolved gas in liquid.
2. The gas density increases with increasing pressure, thus in total the density difference
between gas and liquid decreases.
A common expression for the interfacial tension pressure dependence is given by
σ ( P)
= 1− K ⋅ P
σ0
where σ 0 is the interfacial tension for pure liquid and K is a constant. The behaviour of the
equation is shown in figure 2.6.
Chapter 2 Definitions and basic quantities - Version 2017/1 17

A
qG uG
AL
uL
qL AG

Figure 2.7: Quantities entering the definitions of velocities and fractions

2.2 Flow velocities


Before we proceed to consider the dynamics of two-phase flow, i.e. the influence of forces on
motion, we will define some quantities that describe the flow velocity. We refer the
definitions of quantities to figure 2.7. We use subscript L for liquid and G for gas, unless
otherwise stated.

2.2.1 Volumetric flow rates


Gas and liquid flow with volumetric flowrates qL and qG in the pipe., referring to the volume
transport rate of liquid and gas through any given fixed cross section (“station”) of the pipe.
Alternatively the flow can be defined by the superficial velocities or by the phase velocities.
These are introduced in the next sections.

In some cases it is useful to let the imagined station moves with a velocity (U). This is done in
the “drift flux model” discussed in chapter 5.

2.2.2 Mass flow rates


The units for mass flow rates is kg/s and is the equivalent transport of mass through the
station. Common symbols for mass flow rate is GL,G or m L ,G . In the latter case the dot refers
to time derivative (of mass). Note that there is some inconsistency in the international
literature regarding symbols. One may found papers where m L ,G means flux of mass, i.e. mass
flow rate divide by pipe cross section. Mass flow rate is related to other quantities by the
following relations

GL = ρ L ⋅ q L
GG = ρ G ⋅ q G
G = G L + GG
GG
Often the quantity x is used to define a “dynamical gas fraction”: x = called the mass
G
dryness fraction, or simply the “quality” in connection with steam (nuclear power generation).
18 Chapter 2 Definitions and basic quantities - Version 2017/1

2.2.3 Superficial velocities and mixture velocity


The superficial velocities are defined by
q
U LS = L
A
q
U GS = G
A

They are also referred to as apparent velocities or volumetric fluxes. From the definition we
see that if the volumetric flowrates and the pipe cross section A is known, the superficial
velocities follow directly. We don’t to know more about the flow to determine the superficial
velocities. The physical interpretation is thus simply the volumetric flowrates divided by the
pipe cross section.

However, the sum of the superficial velocities, called the mixture velocity U mix = U LS + U GS
equals the real average velocity in the flow.

2.2.4 Phase velocities


The phase velocities are the real velocities of the flowing phases. They may be defined locally
(at a certain position in the pipe cross section) or as a cross sectional average for the pipe.
They are defined by

qL
uL =
AL
qG
uG =
AG

In order to determine these quantities it is necessary to determine the real flowing cross
sections AL and gas AG for liquid and gas. This is equivalent to knowing the fractions or
amount of liquid and gas in the flow. From a metering point of view many measurement
techniques have been developed to determine the phase velocities.
At first sight it might seem as a trivial matter to measure the phase velocities. In practice
however, there is still no “universal” instrument which may function for all flow regimes
encountered in two-phase flow.

2.2.5 Relative phase velocities and slip


Gas and liquid in general flow with different phase velocities in pipe flow. The relative phase
velocity or the slip velocity is defined by

u S = uG − u L

The slip velocity thus has the same unit as the phase velocities. In addition the slip ratio
u
S= G
uL
is commonly used. Note that the slip ratio is dimensionless.
Chapter 2 Definitions and basic quantities - Version 2017/1 19

It may easily be shown that if the slip ratio is 1 (referred to as noslip) the following relation is
valid

u G = u L = U mix

Exercise: Show the relation above.

A note on terminology: In hydrodynamics the word slip is used if two phases have different velocity at
their interface. In particular in reservoir technique slip is known from “high velocity” in porous media (the
Klinkenberg effect). Low molecular density and high flow speed in the pores leads to loss of the continuum
condition y parallel = 0 (noslip ) at the pore wall.

2.2.6 Measurement of flow velocities


In single phase flow the velocities are measured by means of a variety of instruments,
depending on application and accuracy. For traditional industrial flow is used
• Turbine meters
• Orifice meters
• Acoustical Doppler flowmeters
• Vortex meters

They have inaccuracy at the order 1% . In recent years also

• Coriolis flowmeters

have gained increasing popularity, being accurate to better than 0.1% for single phase flow.
The Coriolis flowmeter measures the fluid density in the same operation. For research
purposes is often used Laser-Doppler Anemometry (LDA) which can measure velocities
ranging from micrometers pr.second to kilometers pr. second, with extreme accuracy and very
locally. (i.e. point measurements). For transportation measurements LDA is a less feasible
technique. The reason for this is the existence of a flow profile, i.e. the velocity varies from
position to position. It may take several hours to obtain the velocity profile in pipe flow by
LDA. A much faster and more recent laser-based technique is the Particle-Imaging
Velocimetry (PIV) technique. This is in principle a fast photographic technique where the
laser is used as flash, and a camera takes pairs of photos that enter a mathematical analysis
called two-dimensional cross correlation.
For two phase flow measurement the most rapidly developing technique is based on using
cross correlation function (CCF) analysis on pairs of sensors along the flow. This technique in
principle detects the movement of fluctuations in the flow. Such fluctuations can be
associated with gas fraction, water concentration or even temperature. The CCF analysis is
capable of measuring the time T it takes for a disturbance to flow between two stations.
Knowing the distance L between the stations the velocity is calculated as L/T. This will be
further discussed in chapter 7.
20 Chapter 2 Definitions and basic quantities - Version 2017/1

VG

vL VL
vG

Local fraction v Average volume fraction V


v
εG = G VG
v εG =
V

Figure 2.8: Definitions of volume fraction

LINE AVERAGE Gamma detector

Gamma source

AREA AVERAGE Capacitance sensor

Electronics

VOLUM AVERAGE

Quick closing valves

Figure 2.9: Techniques for fluid fraction measurements


Chapter 2 Definitions and basic quantities - Version 2017/1 21

2.3 Fluid fractions


With reference to figure 2.8 and previous definitions of velocities we now discuss the most
common fluid fraction definitions. Depending on how phase amounts (“volumes”) are
measured, the phase fractions are always averages over a certain domain.
We discriminate between
• Line average
• Area average
• Volume average
even though in all these cases the average is always taken over some kind of volume.
Figure 2.9 illustrates the three basic averaging measurements. Line averages are measured e.g.
with a narrow beam gamma(γ)-densitometer. Area averages we obtain with a wide beam
gamma-densitometer, or with (electric field) impedance sensors having electrodes made into
the pipe wall. Volume averages (pipeline average) may be obtained by using quick closing
valves. Gas and fluid fractions are then defined by
V A L
ε G = G or = G , or = G for gas, and
V A L
V A L
ε L = L or = L , or = L for liquid (subscript L)
V A L

where V, A and L are volume, area and length respectively. If the flow pattern was a
completely homogenized mixed the three averages would be equal. In two-phase flow
terminology ε G should be referred to as gas fraction, however the term “void fraction” is seen
very often. Gas is then considered as abscence of liquid (i.e. void ). In petroleum industry is
still found the symbol HL instead of ε L , referred to as the liquid holdup. Sometimes this is
also used for the amount or fraction of liquid (oil) in a total pipeline. Finally one may find
another set of symbols, α
for gas fraction and β for liquid fraction. In three phase flow (oil-water-gas) flow β is used
for oil fraction, while γ is used for water fraction.

Exercise: Show that in two-phase gas-liquid flow α + β = 1 . What is the equivalent for three
phase flow. How many independent fractions are there in each case?

2.3.1 Noslip fraction


In many cases we do not know the dynamics in the flow, and it becomes impossible to
measure or calculate the fraction of gas and liquid exactly. Nevertheless it may be important
to estimate the fluid fractions e.g. in a subsea pipeline. Provided we know the volumetric
flowrates qL and qG we may determine the so called noslip fractions (or flux fractions):

qL
λL =
q L + qG
qG
λG =
q L + qG
The reason for the name noslip fractions is obvious from the next section.
22 Chapter 2 Definitions and basic quantities - Version 2017/1

2.3.2 True fractions at slip


If slip is present, i.e. the phase velocities are different , the fluid fractions are different from
the noslip fractions. However it is possible to determine the true fractions if the slip ratio is
known. The true liquid fraction, for a start, is given by
A AL
εL = L =
A AL + AG

We may also consider the phase velocities

qL
uL =
AL
qG
uG =
AG
implying that
qL U LS
εL = =
1 1
q L + ⋅ qG U LS + ⋅ U GS
S S

where S is the slip ratio. Similarly we obtain


qG U GS
εG = =
S ⋅ q L + qG S ⋅ U LS + U GS

In the limit S → 1 we get ε G = λG and ε L = λ L .Thus the noslip fractions become equal to the
true fractions, but only if S=1. This is the reason for the name noslip fractions for λG and λ L .

Exercise:
1. Determine ε L and ε G for large S.
2. Assume that phase mixing increases with increasing mixture velocity. Explain why this
implies that S → 1 . Also assume that phase separation is important at low mixture
velocities. Draw a graph showing how ε L varies from small to large velocities.

2.3.3 Measurement of fluid fractions


As mentioned in the previous sections there are several instrument techniques that are used
for measurement of fluid fractions. The intuitively correct measurement method consists of
grabbing a well known volume from the flow, let it separate at rest maintaining the same
pressure and temperature as in the pipe, and measure the liquid and gas volumes separately.
This is to a certain extent achieved by means of quick closing valves, or by sampling. A
particular technique is called isokinetic sampling. Here the sampling procedure is based on
grabbing or sampling the fluid in such a way that the flow in the pipe is only very little
disturbed by the grabbing equipment.
However, it is not feasible to cover all kind of flows, industrial or research with sampling
techniques. Quick closing techniques, although closing fast, takes at least several minutes and
Chapter 2 Definitions and basic quantities - Version 2017/1 23

maybe hours for the fluids to separate properly. Thus it is a slow technique. Furthermore, the
closing procedure may introduce pressure pulses which are unwanted. And finally, the
identification of liquid level requires either sophisticated instruments or human observation.
This makes the technique useless for industrial flow, and at least for subsea measurement or
measurement downhole petroleum wells.
The most widely used techniques today are those shown in figure 2.9, namely γ (gamma) -
densitometers using various beam widths and gamma energies. The attenuation of the beam is
related to the mass of fluids crossing the beam. A particular version of this technique is use of
X-rays instead of the traditional technique based on using radioactive isotopes (americium,
cesium and cobalt are common). X-ray techniques have been applied as part of a PhD work at
the HiS two-phase flow laboratory (Albrechtsen, Time and Aas, 1995)
The other main group is impedance sensors, which may be divided into resistivity sensors and
capacitance sensors. This will be discussed in more detail in chapter 7. For oil continuous
flow capacitance sensors are most common. This technique is also used widely at the HiS
laboratory and ha been part of a PhD work (Time, 1993).
For water-oil-gas mixtures impedance is used as a combined technique, where both the
resistive and the capacitive components are measured. Christian Michelsen Research in
Bergen applied this technique. Fluenta (Bergen) has commercialised the concept into a
multiphase flowmeter.
Stanford Research in California worked along a parallel technique based on microwave
attenuation, particularly for detection of water in oil. The sensitivity to water in oil is very
high (in the ppm range). This concept was later taken over by Hitec (Forus, Stavanger) now
Multi Fluid International (MFI). In combination with a gamma densitometer it is used as a
three phase flowmeter.

2.4 Two-phase fluid properties – mixing rules


In this section we present calculation procedures for determination of effective density and
viscosity for a two-phase process, provided we know the fluid fractions and the single phase
fluid properties. Such procedures or equations are sometimes referred to as mixing rules.

2.4.1 Density
The density for a two-phase mixture is a well-defined, geometric quantity that can be
calculated provided the fluid fractions are known. The equation

ρ m = ρ Lε L + ρ Gε G

gives an exact result.

2.4.1 Viscosity
Unlike the density, viscosity for a mixture is not a well-defined quantity just in terms of fluid
fractions and single phase viscosities. The mixture viscosity in fact depends strongly on
dynamical processes as well including bubble size, flow regime etc. It is therefore not
unexpected that many different models exist.
24 Chapter 2 Definitions and basic quantities - Version 2017/1

The most common viscosity models are:

Cichitti: µ m = xµ G + (1 − x) µ L

1 x (1 − x)
McAdams: = +
µm µG µL

Dukler: µ m = ε G µ G + (1 − ε G ) µ L

Geometrical average: µ m = µGε ⋅ µ L(1−ε


G G)

Some of these models are illustrated in figure 2.10. Cichittis model is used mainly for small
gas fractions (continuous liquid – dispersed bubble flow) while McAdams model mainly is
used for low liquid fractions (continuous gas – annular mist flow). Finally we introduce the
early model by Einstein (1906) for viscosity of a fluid with dispersed particles having a
volume concentration α of particles.

µm  µ + 52 µ 
= 1+ α 
µ  µ+µ 
 

where µ is the viscosity of the continuous fluid, while µ is the internal viscosity of the
dispersed particles (e.g. bubbles)

Viscosity models
ρG = 100 kg/m3
ρL = 1000 kg/m3
µL

Chichitti
Mixture viscosity

Dukler

McAdams

µG
0 Gas fraction 1.0

Figure 2.10: Viscosity models versus gas fraction


Chapter 2 Definitions and basic quantities - Version 2017/1 25

Example: Assume liquid and gas flow separated and laminar (non turbulent) between two
plane plates. The upper plate moves while the lower is static. Determine the effective
viscosity of steady state flow in the two situations that:
1. The phases flow in parallel between the two plane walls, with the interface parallel to the
walls.
2. The phases flow as sequential segment, i.e. with the interface normal to the walls.
a) Moving plate
Gas U
Liquid
U1
Pressure gradient

b) Moving plate H
L
y
Liquid

Gas

Close-up of a)

Solution: These two cases may be considered as extremes concerning effective viscosity
models. In the first case the flow is a seen on the right figure a). The velocity profile is linear
in each phase. Also the viscous shear stress must be equal in the two phases because of
Newtons 3. Law at the interface.

 du  ∆u U
The shear stress in the liquid is τ L = µ L ⋅   = µ L ⋅ 1 = µ L ⋅ 1
 dy 1 L L
 du  ∆u 2 U − U1
and in the gas τ G = µ G ⋅   = µ G ⋅ = µG ⋅
 dy  2 H −L H −L
L H −L
Then ∆u1 = τ L ⋅ and ∆u 2 = τ G ⋅
µL µG
Also τ G = τ L = τ where τ is the total apparent shear stress given by
 du  U H
τ = µ m ⋅   = µ m ⋅ → U =τ ⋅
 dy  H µm
H L H −L
The velocity of the upper plate is U = ∆u1 + ∆u 2 thus τ⋅ =τL ⋅ +τG ⋅
µm µL µG

1 1 L 1 H −L 1 ε L εG
which leads to = + or: = +
µm µ L H µG H µm µ L µG

Situation b) leads to the result: µm = ε L µ L + ε G µG

The proof is left to the reader.


26 Chapter 2 Definitions and basic quantities - Version 2017/1

2.4.3 Interfacial tension of two-phase flow mixtures


The interfacial tension of a two-phase mixture is an even less well-defined quantity than two-
phase viscosity. There is a strong dependence on how gas is mixed into the liquid surface, as
well as the presence of small and large bubbles. At times, when turbulence becomes strong
there might even be doubt as to where the interface between gas and liquid starts and stops.
Nevertheless one may find expression like the following linear mixing rule, here applied to
the interface of a water (W)-oil(O) mixture facing a gas

σ m = ε W σ WG + ε Oσ O

The equation should never be taken too seriously. One may easily find situations where the
equation is not valid, e.g. in situations where foams or micro emulsions appears. In such cases
the emulsion becomes a new phase which itself may have so strong mechanical properties that
the interfacial tension between the bulk liquid gas phase apparently increases beyond the
value of each of the individual interfacial tensions σ W or σ O .

2.5 Flow regimes


In single phase flow we discriminate between laminar and turbulent flow. In two-phase flow
we discriminate in addition between flow regimes that are characteristic for the time and
space distribution of gas and liquid flow.

2.5.1 Horizontal flow


In horisontal flow we discriminate between the flow regimes
• Stratified flow
• Slug flow Increasing flow
• Dispersed bubble flow velocity
• Annular flow

These are shown in figure 2.11. At low velocities the gas and liquid are separated as in
stratified flow. At high velocities gas and liquid become mixed. Slug flow is an example of a
flow regime in between, representing both separation and mixing. Slug flow is consequently
referred to as an intermittent flow regime.

2.5.1 Vertical flow


In vertical flow we discriminate between the flow regimes
• Slug flow
• Churn flow Increasing flow
• Dispersed bubble flow velocity
• Annular flow

Figure 2.12 illustrate the flow regimes in vertical flow. The same comments that apply to
horizontal flow are valid in vertical flow. The big difference is that in vertical (cocurrent
Chapter 2 Definitions and basic quantities - Version 2017/1 27

upward) flow it is not possible to obtain stratified flow. The equivalent flow regime at
identical flowrates of gas and liquid is slug flow with very slow bullet shaped Taylor bubbles.

Discussion : Is it possible to have stratified flow regime in vertical downward flow?


28 Chapter 2 Definitions and basic quantities - Version 2017/1

Flow direction
Stratified smooth
flow (SS)
Low velocity

Stratified wavy
flow (SW)
Intermediate velocity

Elongated bubble
flow (EB)

Slug flow (I)

Dispersed bubble
High velocity

flow (DB)

Annular (wavy)
flow (A-AW)

Dispersed bubble (DB)


10
Taitel and Dukler
2.5 cm i.d. Pipe
air-water flow
at atmospheric
Elongated Slug flow (I) conditions
1 bubble flow
(EB/I) Mandhane
ULS (m/s)

Annular -
annular wavy -
annular mist flow
0.1
Stratified
wavy
Stratified smooth flow
flow (SS) (SW)

0.01

0.1 1 10 100
UGS (m/s)

Figure 2.11: Flow regimes and flow regime map in horizontal two phase flow.
Chapter 2 Definitions and basic quantities - Version 2017/1 29

Dispersed bubble

Annular (wavy)
Flow direction

Churn Flow
Slug flow (I)

flow (A-AW)
flow (DB)

Low velocity Intermediate velocity High velocity

10
DISPERSED
BUBBLE Theoretical flow pattern map
(small) ( Barnea, Taitel and Dukler )

Upward vertical air-water flow


ANNULAR

1
CHURN

DISPERSED D = 5 cm ( i.d. )
P = 1bar, T=25 0C
BUBBLE
ULS [m/s]

lE/D =200
(medium)
0.1

SLUG

0.01

0.001
0.01 0.1 1 10 100

UGS [m/s]

Figure 2.12: Flow regimes and flow regime map in vertical two phase flow.
30 Chapter 2 Definitions and basic quantities - Version 2017/1

2.6 Friction factor, shear stress and pressure gradient


The pressure gradient dP/dx in pipe flow depends on:
Pipe diameter D, fluid viscosity µ , fluid density ρ and flow velocity U

In addition the wall roughness and pipe inclination is important. In multiphase flow the flow
regime is also important. In single phase and multiphase flow the discrimination between
laminar and turbulent flow plays a decisive role for the friction pressure drop. The type of
flow is determined from the Reynolds number

ρUD UD µ
Re = ≡ where we have introduced the kinematic viscosity ν =
µ ν ρ

We discriminate between the following regimes:

• Re ≤ 2000: Laminar flow


• 2000 < Re ≤ 4000: Transition between laminar and turbulent flow
• 4000 < Re: Turbulent flow

Completely turbulent flow (“homogeneous isotropic turbulence”) is achieved only at very


high Reynolds numbers, Re ≈ 104 – 105. The total pressure gradient in the pipe may be
considered as composed of 3 different terms:
 dp   dp 
Frictional pressure gradient   , hydrostatic pressure gradient   and acceleration
 dx  f  dx  h
 dp 
pressure gradient   . Thus
 dx  a

dp  dp   dp   dp 
=  +  + 
dx  dx  f  dx  h  dx  a

The contribution from each of these are different in single phase and two phase flow.

2.6.1 Single phase flow


Frictional pressure drop – Fanning friction factor
In single phase laminar flow one may show that the frictional pressure gradient at constant
flow velocity and constant pipe diameter is given by

 dp  4 16 1 16
 = ⋅ ⋅ ρ U2 with laminar friction factor f ≡ (Fanning friction factor)
 dx  D 
Re 2 Re
f

Type C n
Blasius 0.079 0.25
Dukler 0.046 0.2

Table 2.2 Values for the turbulent friction factor ( Blasius form, Fanning type )
Chapter 2 Definitions and basic quantities - Version 2017/1 31

4
This equation form, containing the term is based on friction factor of the Fanning type.
D
We will soon meet the alternative Moody type friction factor which simply “absorbs” the
number 4 into the friction factor. In laminar flow the friction factor is exact and can be
calculated theoretically, as a result of the well defined parabolic velocity profile.

In turbulent flow the friction becomes larger. This is due to the velocity profile becomes more
uniform (although fluctuating) causing a larger velocity fall-off towards the pipe wall and thus
a larger shear. In effect it may seem as if the viscosity apparently increases. This implies that
a new friction factor is required. For high Reynolds numbers we use the form
f = C ⋅ Re − n

which is often referred to as the (powerlaw) Blasius form. The most common values for C and
n are given in table 2.2 In this course the Dukler values are mostly used. Note that also the
laminar friction factor fits into the Blasius form.

Moodys friction factor


In some connections we find the Moody friction factor occurring as

 dp  1 64 1 64
 = ⋅ ⋅ ρ U2 which for laminar flow is f M ≡
 
dx D  2
Re Re
fM

Note the difference both in the frictional pressure drop equation and in the friction factor.

Commonly used turbulent friction factors


We present some well known turbulent friction factors.

1. Drew, Koo and McAdams (1932) smooth pipe friction factor:

f = 0.0056 + 0.5 Re −0.32


This correlation is valid for Reynolds numbers in the range 3000 < Re < 3 .106.

2. Nikuradse rough pipe friction factor:


1  2ε 
. − 2 log 10  
= 174
f  D

This formula is based on relative roughness ε D .Note that ε itself has dimension length.
The way it is presented it is of Moody type to fit the graphical data of the Moody plots.

3. Colebrook & White (1939) friction factor for rough pipes:

1  2ε 18.7 
= 1.74 − 2 log10  + 
 
f  D Re f 
32 Chapter 2 Definitions and basic quantities - Version 2017/1

This is Nikuradses friction factor combined with an equation for friction factor in smooth
pipes as discussed in chapter 3. This equation is implicit in f and has to be solved
numerically. This can be done by rewriting the equation into
2
 
 
 1 
fu =  
  ε 
1. 74 − 2 log  2 + 18.7 
 10
 D Re f 
  i 

Her fi is an initial guess, entered into the equation and fu may be calculated. Afterwards the
replacement fi = fu is done, and a new iteration carried out. The procedure stops when
iteration k
fi − fu
< accepted error (AE)
fi

The procedure is normally quite fast and converges for AE = 0.001 after a few iteration.
Alternatively f may be determined from Moody graphs, or by an explicit formula by
Haaland (1983):
1  ε / D 1.11 6.9 
≈ −1.8 ⋅ log10  ( ) + 
f  3.7 Re 

Material ε (ft)
Glass 10-6
Drawn tubing (extruded steel) 5.10-6
Commercial steel 1.55.10-4
Asphalted cast iron 4.10-4
Cast iron 8.5.10-4
Galvanized iron, new 5.10-4
3 years old 9.10-4
Concrete 10-3 - 10-2
Wood stave 6.10-4 - 3.10-3

Table 2.3: Roughness in pipes of various materials.

Matlab example: Use Matlab to generate curves for Colebrook-White friction factor as a
function of Reynolds number.
Solution: Define Reynolds numbers ranging from 103 to 108. For each value of the Reynolds
number, calculate the friction factor for given relative roughness. The Matlab loop just to
calculate the friction factor according to the iterative solution is given below. The array x
contains the 100 Reynoldsnumber values.
Chapter 2 Definitions and basic quantities - Version 2017/1 33

for i=1:100; % Loop over Reynolds number (here called x)


if x(i)<4000; % Laminar flow
f(i)=64/x(i);
ff(i)=f(i);
else % Colebrook & White method (1939) - iterative solution:
fgam=1;
feil=1000;
while feil >= 0.01
fny=1/(1.74 - 2*log10(2*eps(j)+18.7/(x(i)*sqrt(fgam))));
fny=fny*fny;
feil=abs((fny-fgam)/fgam);
fgam=fny;
end % end of the while-test
f(i)=fny;
end

end % end of Reynolds number for loop i

The complete program (can be obtained as part of the course material) gives the following
plot:
-1
10

Relative
roughness
Moody friction factor (fM )

10-2

10-3
Laminar
flow
10-4
-2
10
10-5

3 4 5 6 7 8
10 10 10 10 10 10
Reynolds number (Re)
ρUD
Reynolds number Re =
µ

Figure 2.13: Moody diagram


34 Chapter 2 Definitions and basic quantities - Version 2017/1

Wall shear stress and frictional pressure drop


The friction factor is related to the wall shear stress by the equation
1
τW = f ρ U2
2

This equation is valid both in laminar and turbulent flow. In terms of wall shear stress the
frictional pressure drop may be written as:
 dp  4
  = ⋅τ W (Fanning)
 dx  D

Hydrostatic pressure gradient


The hydrostatic pressure gradient in a fluid is given

∇Ph ( y ) = ρ g iy

where Ph ( y ) is the pressure in still fluid as a function of the depth y, and i y is the unit vector
in the direction of gravity (downwards). The component of the hydrostatic gradient in a pipe
is often called the hydrostatic pressure gradient in the pipe. With a pipe inclination β relative
to the vertical direction, the hydrostatic pressure gradient is given by
 dp 
  = ρ g cosβ
 dx  h

Acceleration pressure gradient


In stationary single phase flow the pressure changes if flow velocity U changes due to pipe
expansion or contraction. In one-dimensional description the pressure gradient is given by

 dp  dU
  = −ρ U ⋅ (Note the sign)
 dx  a dx

As we will see in chapter 3 the acceleration pressure gradient is a natural consequence of


Bernoullis equation. Note that the acceleration pressure gradient is not dissipative, i.e. total
energy is conserved, although pressure changes. The balancing energy term is the kinetic
energy 12 ρU 2 .

Exercise: Discuss positive and negative sign of the various pressure gradients, by answering
the following questions:

Claim True False


P drops in the direction of flow if only friction is present
P increases in the upward direction in a static fluid
P increases when liquid flows from a wide pipe into a narrow pipe
P decreases when liquid flows upward in a pipe which becomes narrower in
the flow direction
Chapter 2 Definitions and basic quantities - Version 2017/1 35

2.6.2 Calculation of pressure drop in two phase flow


For two-phase flow calculations we will start with the assumption that the fluids mix to a
homogeneous new fluid with properties (density and viscosity) given by the mixing rules
given previously in section 2.4. This approach is called the homogeneous two-phase pressure
drop model. Calculation may be done by assuming either no-slip (S = 1) or by specifying a
certain slip ratio in order to obtain appropriate fluid fractions. The Beggs and Brill pressure
drop model that is introduced in chapter 6 applies a homogeneous model for friction and
hydrostatic pressure, some parts calculated with the no-slip assumption, and some taking into
account the real slip.

Frictional pressure drop


As in single phase we use

 dp 
  = ⋅ C (Re m ) ⋅ ρ m U mix
4 −n 1 2

 f
dx D 2

where index m means “mixture”. The homogeneous model is realistic essentially in turbulent
well mixed flow, which means the selection, C=0.046, n=0.2 should be used.
The mixture Reynolds number is calculated as
ρ U D
Re m = m mix
µm
and the fluid properties are determined as in section 2.4

Hydrostatic pressure gradient


The hydrostatic pressure gradient is calculated as
 dp 
  = ρ m g cosβ (angle β relative to vertical direction)
 dx  h
where only the mixture density is required compared to single phase flow.

Acceleration pressure drop


The two most important contributions come from
• Change in pipe cross section and velocity change
• Change in gas density
We will not go further into this gradient associated with gas density here, since it will be
discussed in connection with the Beggs and Brill model. The acceleration pressure gradient
also is analogous to in single phase flow
 dp  dU mix
  = − ρ m U mix ⋅
 dx  a dx
modifying only the density and using the mixture velocity.
36 Chapter 2 Definitions and basic quantities - Version 2017/1

2.6.3 Calculation of pressure traverses in long pipelines

In long pipelines the pressure drops considerably from inlet to outlet, thus the gas density and
thereby the gas velocity change accordingly during the passage. In addition, in real
hydrocarbon transport the evaporation of gas from the oil contributes to increase the gas flow
velocity. This is to some extent a self amplifying effect, since higher velocity implies higher
pressure drop and thus even higher evaporation. The pressures at inlet and outlet thus in
reality determine the total flow velocity. Factors which further complicate the calculations are
the temperature profile along the pipeline and the heat conduction from surroundings into the
pipeline. If the pressure gradient along the pipe is known (based on velocities, pipe geometry
and fluid properties) we may find the total pressure drop over a pipeline with length L as:

L
dp
∆P = ∫ ( x ,U LS ,U GS , α ) dx
0
dx
where α represents all other parameters than the flow rates. However, the volumetric flow
rates (qG essentially) are pressure dependent and the calculation must be carried out iteratively
because the pressure profile is not known. Numerically we solve the problem by dividing the
pipe into segments. The problem is solved by stepwise calculation of pressure along the pipe
until pressure drop and flowrates match each other. Two different problems are normally
solved, depending on whether pressure drop or velocities are known.

1. Assume the inflow of gas and oil is known by the flowrates qL and qG. Also the pressure at
the inlet is assumed known. Calculate the pressure along the pipeline including the outlet
pressure.
This is the simplest case and may be found by piecewise solution along the pipeline. The
pressure P is calculated successively along the pipeline from inlet to outlet. The gas
flowrate is updated for each cell as a function of P. For each discrete cell the pressure
gradient is calculated and multiplied with the cell length ∆x to obtain the pressure in the
next cell. The outlet is the end of the last cell.

2. Assume the pressure at inlet and outlet are known ( Pin∗ and Pout

) and that the flowrates qL
and qG are unknown. Calculate the inlet flowrates that are responsible for the pressure
drop.
This problem has to be solved iteratively. Start values for the flowrates are initially
1
guessed. Then the solution follows the procedure in 1, and the outlet pressure Pout is

found. This pressure is compared with the imposed outlet pressure Pout .
1 ∗
If Pout differs from Pout the error

∆Poutj = Poutj − Pout
is used to make improved estimates of inlet pressure. The iteration continues until the
absolute value of the error becomes smaller than a given tolerance.

It is possible to solve the problem by other direct numerical methods, e.g. by matrix inversion
as in one dimensional reservoir simulation. The problem here is slightly worse, because the
equations are more non linear than in Darcy flow. Linearization of the equations simplifies the
calculation, but removes many important fluid dynamical effects. The solutions thus may
become erratic and eventually unstable.
Chapter 2 Definitions and basic quantities - Version 2017/1 37

Matlab example: Write a Matlab program that calculates the pressure variation along a 5000
meter high closed gas well, containing an ideal gas with density 0.7 kg/m3 at reference
pressure 1 atm. Compare with the analytical solution based on ideal gas equation of state, as
well as with the solution based on assuming constant gas density everywhere in the well.

Solution : The program with comments and selected vaules of variables is given below.
Copy and run it in Matlab to see the results.

% This program calculates the pressure variation in a closed-in gas well.


% *******************************************************************************
clear
clf
P0 = 110*10^5 % Pressure at top of gas well (Pa)
rho_ref = 0.7 % Gas density at ref.pressure (kg/m^3)
P_ref = 1.013*10^5 % Reference pressure for ideal gas (Pa)
g = 9.81 % Gravitational acceleration m/s^2

dH = 100 % Height interval


H = 0:dH:5000 % Height array
P = P0*exp(rho_ref*g*H/P_ref) % Analytical solution for pressure at well bottom
P2 = P0 + (rho_ref*P0/P_ref)*g*H % Bottom pressure if gas denity was constant from top to bottom
PP(1) = P0 % Start pressure for stepwise simulation

NH = length (H) % Determine array lenth of H

for i=2:NH % Loop over well segments


dens = rho_ref*PP(i-1)/P_ref; % Calculate gas density in each well segment
dP = dens*g*dH; % Calculate pressure increase over segment
PP(i)=PP(i-1)+dP; % Update pressure for next segment top
end

figure(1) % Plott pressures


plot(H,P,'-ok', H,P2,'-*g', H,PP,'-+r')
legend('Analytic','Constant density','Stepwise solution', 0)
xlabel('Well Depth (m)')
ylabel('Pressure (Pa)')
xp=1000
yp=1.5e7
err = (P(NH)-PP(NH))/1.e5

tekst = num2str(err)
stepp = num2str(dH)
text(xp,yp,['Error = P(Analytic) - PP(Stepwise) = ' tekst ' bar'])
text(xp,yp-5e5 ,['Step length along well = ' stepp ' m'])
38 Chapter 2 Definitions and basic quantities - Version 2017/1

7
x 10
1.55

1.5 Error = P(Analytic) - PP(Stepwise) = 0.17646 bar

1.45 Step length along well = 100 m

1.4
Pressure (Pa)

1.35

1.3

1.25

1.2
Analytic
1.15 Constant density
Stepwise solution
1.1
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Well Depth (m)

Figure 2.14: Pressure profile in 5000 meter deep gas well, based on ideal gas equation of
state.

2.7 Summary
We have introduced the basic definitions of

• Fluid fractions (mass and volume based)


• Velocities (phase velocities and superficial velocities)
• Slip

for two phase flow. The fluid properties for single phase flow have just barely been discussed
since they enter in other courses in petroleum technology. The interfacial tension is an
important quantity for two phase flow defined early in this chapter.

Fluid properties for two-phase flow systems may be determined from single phase properties
by means of mixing rules where the fluid fractions are used as weight factors.

This chapter also presented the flow regimes in vertical and horizontal flow. Pressure drop
models were adopted from single phase flow and by means of the homogeneous model two
phase flow pressure drop can be calculated without knowledge about flowregime.
Chapter 2 Definitions and basic quantities - Version 2017/1 39

2.8 Exercises
1. Show Laplace’s equation by using the method of virtual work.

2. Determine the constant K in the expression for surface tension versus pressure for the
various gases in figure 2.6.

3. Oil having viscosity 3 cP and density 850 kg m-3 flows in a smooth pipe with inner
diameter d = 15cm. The flow velocity is U = 4 m/s. Calculate the friction factor by means
of the correlation by Drew, Colebrook and White, and the Blasius formula. Compare the
calculations with the Fanning factor used by Taitel and Dukler.

4. When 0.0019 m3s-1 water flows in a d = 76 mm pipeline, is the flow then laminar or
turbulent? The water viscosity 1 mPa.s (milliPascal seconds).

5. The liquid fraction was measured to 0.35 in a 8'' pipe where liquid and gas flow with flow
rates 0.5 ft3/s and 3.25 ft3/s respectively. Calculate superficial liquid and gas velocity, true
phase velocities, slip velocity and no-slip liquid fraction.

6. Assume oil and gas flow upward in a pipe with inclination 450. The pipe diameter
gradually increases from D1 = 5 cm at the bottom, to D2 = 15 cm at the top. The pipe
length is 100 m. The mass flow rate of oil is 2.0 kg/s and for gas 0.2 kg/s. The oil density
is constant, equal to 900 kg/m3. The gas is ideal with molar weight M = 25. The pressure
at the lower (inlet) end of the pipe is 90 bar and the temperature is T = 15 0C. The
viscosity of oil is 2 cP and for gas equal to 0.03 cP. Assume noslip between gas and
liquid.

a. Can the flow regime be stratified? Give a short qualitative explanation.


b. Calculate the liquid fraction at the bottom of the pipe.
c. Calculate the contribution to the pressure gradient at the bottom of the pipe from
• Flow acceleration
• Change in potential energy (hydrostatic gradient)
• Friction

7. The interfacial tension in an air-water system is σ = 72 mN/m.


a. Determine the pressure between the inside and outside of an air bubble with radius
1.) r = 1 cm
2.) r = 0.01 mm
b. A capillary tube with radius r = 0.01 mm stands perpendicular into a bowl of water.
Determine the level over water in the pipe?
c. Draw a figure showing the pressure variation in the tube from bottom to top across the
gas-liquid interface.

8. Show the equations for liquid fraction involving both slip and noslip.

9. Compare Einsteins viscosity model with the models of Cichitti and McAdams for small
fractions of the dispersed phase.
40 Chapter 2 Definitions and basic quantities - Version 2017/1

10. Which type of friction factor is the Moody diagram based on?

11. Determine the Reynolds number where Duklers friction factor equals the expression for
the laminar friction factor. Is the friction continuous at the transition to turbulent flow?

12. Show that the true gas fraction ε G can be expressed by the mass dryness fraction x. Derive
the expression.

Matlab problems

13. Modify the Matlab program for the closed-in gas well to study the pressure variation in
the well if the step length is increased from 100 to 1000 m.

Units conversions:
1 ft = 0.3048 m
1 inch ('') = 2.54 cm
1 cP = 1 mPa s.

You might also like