You are on page 1of 77

July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.

5in b1881-set-I-v1-ch04 page 23

Chapter 4
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

Hydrodynamic Models Based on Flow Patterns


by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

In the previous chapter, simplified methods for the calculation of pressure drop
and void fraction were presented. Improved methods for calculating hydrodynamic
parameters as well as heat and mass-transfer in two-phase flow, require to consider
the detailed structure related to the interfacial distribution of the two phases in
the pipe.
In gas–liquid flow the two phases can be distributed in the conduit in many
configurations called flow regimes or flow patterns, differing from each other in the
spatial distribution of the interfaces. The flow pattern depends on the operational
variables, physical properties of the fluids, and geometrical variables of the system.
Hydrodynamics of the flow, as well as the mechanisms of momentum, heat and
mass-transfer, change significantly from one flow regime to another. As a result,
understanding of the processes in two-phase flow depends on the knowledge of the
existing flow patterns. Determining the flow patterns is the first step for developing
techniques to accurately predict pressure drop, heat and mass-transfer, gas and liquid
holdups, etc.
The main task in classifying the flow patterns is to group together flow
configurations that have common character according to some definitions. The
designation of the various flow patterns should be based on the flow configuration
that has basically the same character, pertaining to the distribution of the interfaces
and the mechanisms dominating pressure drop, heat and mass-transfer. This task
is neither easy nor unique. It depends largely on the individual interpretation of
the researchers. Some researchers detailed as many basic configurations as possible
while others tried to group the flow patterns to the minimum essentials. We chose
to identify as few classes as possible and our suggestion is to identify four major
types: stratified flow, intermittent flow, annular flow and bubble flow (see Fig. 1).
Stratified flow. In stratified flow, liquid flows along the bottom of the pipe with
gas at the top. Both phases are continuous in the axial direction. Stratified flow is

23
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 24

24 Modeling of Gas Liquid Flow in Pipes


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 1. Flow patterns in gas–liquid flow in pipes.

subdivided into stratified smooth, where the liquid surface is smooth and stratified
wavy, where the interface is wavy.
Intermittent flow. In the intermittent pattern, the inventory of liquid in the pipe
is distributed non-uniformly in the axial direction. In horizontal and inclined flow,
plugs or slugs of liquid, filling the whole cross-section of the pipe, are separated by
gas zones that overlay a stratified liquid layer flowing along the bottom of the pipe.
The liquid slugs may be aerated with small gas bubbles. In vertical flow, most of the
gas is located in large bullet-shaped bubbles termed “Taylor bubbles” which span
most of the pipe cross-section. The Taylor bubbles are separated by liquid slugs. The
liquid confined between the Taylor bubble and the pipe wall flows around the bubble
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 25

Hydrodynamic Models Based on Flow Patterns 25

as a thin film. The intermittent pattern is subdivided into slug, elongated bubble and
churn patterns, but the distinction between them has not been clearly defined. The
elongated bubble pattern is considered the limiting case of slug flow at low gas flow
rates, when the liquid slug is free of entrained gas bubbles. Churn flow, on the other
hand is a chaotic flow for high gas flow rates.
Annular flow. In annular flow, the liquid flows as a continuous film around
the perimeter of the pipe. The liquid film surrounds a core of high velocity gas that
usually contains entrained drops. The gas–liquid interface is highly agitated. In the
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

vertical case, the flow is axially symmetric while for horizontal or inclined flow, the
film at the bottom is usually thicker than at the top.
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Bubble flow. In bubble flow, the gas phase is distributed as discrete bubbles
in a continuous liquid phase. In vertical (or near vertical) bubble flow, the gas
phase is distributed more or less uniformly. For horizontal and inclined pipes, the
concentration of the bubbles is higher near the top of the pipe. As the flow rates
increase the bubbles are dispersed more uniformly.
In the following sections, models for the prediction of the hydrodynamic
parameter of each of the afore-mentioned flow patterns are presented. The approach
taken is based on “mechanistic modeling” where the most important physical
processes that affect the behavior of the system are identified and are formulated
mathematically. The predicted results which are based on theoretical mechanistic
modeling, verified by a set of experimental data is, at present, the most practical
approach to solve two-phase flow problems.

1. Steady Stratified Flow


Stratified flow is dominated by the gravity force that causes the liquid to flow as a
continuous layer at the bottom of the pipe. This flow pattern is observed in horizontal
and slightly inclined pipelines. At low gas velocity the interface is smooth. With
increasing gas velocity waves are generated on the gas liquid interface.
The present analysis is a simplified one-dimensional analysis. The geometry of
steady stratified flow in a pipe is shown schematically in Fig. 2. The average liquid

Figure 2. Schematic configuration of stratified flow.


July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 26

26 Modeling of Gas Liquid Flow in Pipes

and gas velocities are UL and UG respectively (the overbars used before for average
quantities are omitted). For prescribed liquid and gas flow rates, physical properties
of the phases, pipe diameter and pipe inclination, the liquid level at steady state and
the pressure drop are calculated.
The momentum equation for the liquid layer in the pipe yields:
dP
− AL − τL SL + τi Si − ρL AL g sin β = 0 (1)
dx
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

where AL is the cross-sectional area of the liquid, SL is the wetted perimeter, Si


is the interfacial perimeter, τL and τi are the shear stresses at the pipe wall and at
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

the interface, respectively, ρL is the liquid density and β is the inclination angle,
considered positive for upflow.
Likewise, the momentum equation for the gas is:
dP
− AG − τG SG − τi Si − ρG AG g sin β = 0 (2)
dx
Considering equal pressure drop in the liquid and the gas, Eqs. (1) and (2) yield
 
τG SG τL SL 1 1
− + τi Si + − (ρL − ρG )g sin β = 0 (3)
AG AL AL AG
The shear stresses are correlated in terms of the friction factors f L , f G and f i ,

ρL UL2 ρG UG2 ρG (UG − UL )|UG − UL |


τ = fL , τG = f G , τi = f i (4)
2 2 2
For a smooth pipe, the friction factors are correlated in the form:
   
DL UL −n DG UG −m 4AL 4AG
f L = CL , f G = CG , DL = , DG =
νL νG SL SG + Si
(5)

where CL and CG equals 0.046 for turbulent flow and 16 for laminar flow. n and m
equal 0.2 for turbulent flow and 1.0 for laminar flow.
The friction factor at the interface is more complex and it depends strongly on
the structure of the interface. For smooth interfaces, it is reasonable to assume that
the friction factor is identical to the smooth solid surface friction factor. Once the
interface is wavy, the interfacial friction factor can be 10-folds and more than for
the smooth case (Cohen and Hanratty, 1968; Andritsos and Hanratty, 1987a). At
this point, we assume that the interface is smooth and that the velocity of the gas is
much higher than that of the liquid; thus,

fi ∼
= f G and UG  UL (6)
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 27

Hydrodynamic Models Based on Flow Patterns 27

For given flow rates, ULS and UGS , Eq. (3) is an implicit equation for the liquid
level h L . Once h L is obtained, the void fraction α = AG / A and the pressure drop
are calculated based on Eqs. (1) or (2).
In order to obtain a generalized solution, the variables are normalized using D
for length, D 2 for area, ULS and UGS for the liquid and gas velocities.
DL SL AL hL UL UG
D̃L = , S̃L = , ÃL = , h̃ L = , ŨL = , ŨG =
D D D2 D ULS UGS
(7)
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

Substituting in Eq. (3) yields


by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

    
−n 2 S̃L −m 2 S̃G S̃i S̃i
X (ŨL D̃L ) ŨL
2
− (ŨG D̃G ) ŨG + + + 4Y = 0
ÃL ÃG ÃL ÃG
(8)
where
 −n
4C L ULS D ρL ULS
2
dP

D νL 2
X =
2
 −m = dP

dx LS
(9)
4C G UGS D ρG UGS
2
dx GS
D νG 2

(ρL − ρG )g sin β (ρL − ρG )g sin β


Y =  −m =
dP
(10)
4C G UGS D
2 ρG UGS

dx GS
D νG 2

X and Y are dimensionless known parameters. X is the well-known Lockhart–


Martinelli parameter and Y can be termed the inclination parameter. The geometrical
dimensionless parameters in Eq. (8) are:

−1
ÃL = 0.25 π − cos (2h̃ L − 1) + (2h̃ L − 1) 1 − (2h̃ L − 1) 2


−1
ÃG = 0.25 cos (2h̃ L − 1) − (2h̃ L − 1) 1 − (2h̃ L − 1)2

S̃L = π − cos−1 (2h̃ L − 1)


(11)
S̃G = cos−1 (2h̃ L − 1)

S̃i = 1 − (2h̃ L − 1)2
à Ã
ŨL = , ŨG =
ÃL ÃG
Solution of Eq. (8) yields
hL
h̃ L = = f (X, Y ) (12)
D
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 28

28 Modeling of Gas Liquid Flow in Pipes

The void fraction and the pressure drop are


AG ÃG 4
α= = = ÃG (13)
A Ã π
and
dP
1 S̃G + S̃i 2 ρG g sin β
 dPdx = (ŨG D̃G )−m ŨG +  dP (14)
dx GS
4 ÃG − dx GS
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

The solution for h L /D is shown in Fig. 3. Four solutions are obtained, depending on
whether the flow combination of the liquid and the gas are laminar or turbulent.
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 3. Solution for the liquid level, horizontal pipe.

Figure 4 includes the solutions for inclined flow (the turbulent–turbulent case
only). Obviously, as the angle of inclination increases the liquid level increases

1/2
(dP/dx)LS (ρL − ρG )g sin β
X= Y =
(dP/dx)GS |(dP/dx)GS |
Figure 4. Equilibrium liquid level for stratified flow (turbulent liquid, turbulent gas).
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 29

Hydrodynamic Models Based on Flow Patterns 29

too, and vise versa. There is a range where multiple solutions are obtained and the
question is what the actual physical solution is? This question is discussed later
when stability of stratified flow is considered.
The results of the calculation for stratified flow are compared to experimental
data and to other models as shown in Figs. 5 and 6. In Fig. 5, the liquid holdup
R(= 1 − α) is plotted versus the Lockhart–Martinelli parameter. As can be seen,
the present theory yields quite accurate results compared to the experimental data
and it is somewhat better than the other models.
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

The pressure drop is also predicted very well as shown in Fig. 6. Note that
the Lockhart–Martinelli method predicts pressure drop which is about twice the
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

measured pressure drop. It is because the Lockhart–Martinelli correlation is based

Figure 5. Liquid holdup in horizontal stratified flow.

Figure 6. Pressure drop in horizontal stratified flow.


July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 30

30 Modeling of Gas Liquid Flow in Pipes

on data taken at relatively high gas and liquid flow rates and not for stratified flow
conditions.

2. Steady Annular Flow


A schematic description of annular flow is shown in Fig. 7. A liquid film of thickness
δ flows along the wall with an average velocity, ULF . For horizontal and inclined
pipes, the film is usually thicker at the bottom, while for vertical and off-vertical
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

flows and for high gas flow rates the film thickness is usually uniform around the
pipe periphery. For simplicity, the film thickness is treated here as uniform.
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

The gas flows as a core in the center of the pipe with a velocity UG . Liquid
may be entrained from the film into the gas core to form liquid drops that are
carried in the gas core with a velocity ULE . The gas core diameter is designated
as DC . Annular flow is basically a separated flow type, as is stratified flow. Yet,
it is more complex owing to the entrainment–deposition process and the existence
of drops within the gas core. In the analysis of annular flow, we may be interested
in the following parameters: pressure drop, film thickness, void fraction, wall and
interfacial shear stresses, fraction of liquid entrained into the gas core, entrainment
and deposition rates.
The formulation of annular flow is similar to that of stratified flow with few
exceptions. The geometry is considered axisymmetric, the gas core is a mixture of
gas and liquid drops and the interface is quite wavy, requiring special knowledge of
the interfacial shear stress.
The momentum equations for the liquid film and the gas core are

dP
− ALF − τL SL + τi Si − ρL ALF g sin β = 0 (15)
dx
and
dP
− AC − τi Si − ρC AC g sin β = 0 (16)
dx

Figure 7. Geometry of annular flow.


July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 31

Hydrodynamic Models Based on Flow Patterns 31

where the subscript LF stands for the liquid film and C for the gas core. ρC is
calculated by
ρC = αC ρG + (1 − αC ) ρL (17)
Equating the pressure drop in the liquid film [Eq. (15)] and the gas core [Eq. (16)]
yields,
 
τL SL 1 1
− + τi Si + − (ρL − ρC )g sin β = 0 (18)
ALF ALF AC
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

where
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

ρL ULF
2
ρC (UG − ULF ) |UG − ULF |
τL = f L τi = f i (19)
2 2
f L is treated as before, whereas f i will be considered later.
 
DL ULF −n
f L = CL (20)
νL
where DL is the hydraulic diameter of the liquid film, DL = 4ALF /SL . The liquid
cross-sectional area is composed of the liquid in the film and the liquid entrained in
the gas core.
AL = ALF + ALE (21)
Since the liquid film is thin, the cross-sectional area of the liquid film is about
ALF ∼
= πDδ.
The liquid flow rate in the gas core is given in terms of a factor FE , which is a
measure of the fraction of liquid that flows in the gas core. Thus,
Q LE = FE Q L (22)
In order to proceed, one needs the velocity of the drops in the gas core. As an
approximation, it is assumed that the liquid drop velocity is very close to the gas
velocity; thus,
Q LE = ULE ALE ∼
= UG ALE (23)
The core void fraction is
ALE FE Q L /UG FE ULS /UG
αC = 1 − =1− =1− (24)
AC A − πDδ 1 − D4 δ
The average void fraction in the pipe cross-section is calculated based on the liquid
flow rate
   
ALE ALF UGS 4δ
FE ULS = UG = UG 1 − α − = 1−α− (25)
A A α D
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 32

32 Modeling of Gas Liquid Flow in Pipes


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 8. Correlation for equilibrium entrainment (Wallis, 1969).

yielding the average void fraction

1− 4δ
α= D
(26)
1+ FE ULS
UGS

The entrainment rate is given empirically, as shown in Fig. 8, it can be fitted by


 1/2
−0.125(φ−1.5) 4 UGS µG ρG
FE = 1 − e φ = 10 (27)
σ ρL

For high gas flow rates, the entrainment rate is very high as most of the liquid flows
in the gas core. For low gas flow rates, φ < 1.5, the core is free of entrained bubbles.

2.1. Interfacial shear


Owing to the wavy character of the liquid film and the entrainment-deposition
process, the interfacial shear is primarily presented by empirical correlations. There
are several ways to define the interfacial friction factor (Hewitt and Hall-Taylor,
1970). The friction factor can be defined on the basis of the relative velocity (UG –
ULF ), the gas velocity UG or the superficial gas velocity UGS and on the basis of the
mean core density ρC or the gas density ρG :
τi
f i,1 = (28)
1
ρ (UG
2 C
− ULF )2
τi
f i,2 = (29)
1
ρ U2
2 C G
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 33

Hydrodynamic Models Based on Flow Patterns 33

τi
f i,3 = (30)
1
ρ U2
2 C GS
τi
f i,4 = (31)
1
ρ (UG
2 G
− ULF )2
τi
f i,5 = 1 (32)
ρ U2
2 G G

and
τi
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

f i,6 = (33)
1
ρ U2
2 G GS
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Note however, that for annular flow usually UG  ULF and UG ∼ = UGS , thus
neglecting ULF and taking UGS instead of UG yields almost the same values for the
friction factor in Eqs. (31)–(33).
Wallis (1969) proposed the following correlation for the interfacial shear stress
in annular flow:

δ
f i,5 = 0.005 1 + 300 = 0.005[1 + 75(1 − α)] (34)
D
The interfacial shear depends on the film thickness δ. This is logical since the
thicker the film is, the higher are the waves on the film which cause an increase
of the friction factor. For a very thin film thickness, the interfacial friction factor
approaches a value of 0.005. For the limit of zero film thickness, the value of 0.005
in Eq. (34) can be replaced by f G = 0.079(ReG )−1/4 .
For the interfacial shear stress coefficient defined in terms of the superficial
velocity, Wallis (1969) proposed
f i,6 = 0.005[1 + 90(1 − α)] (35)
Some other correlations for f i,6 were suggested by Henstock and Hanratty (1976);
for concurrent upward flow:

f i,6
= 1 + 1400F (36)
fs
for concurrent downward flow:
 
f i,6 (1 + 1400F )3/2
= 1 + 1400F 1 − exp − (37)
fs 13.2FG
and for horizontal flow:
f i,6
= 1 + 850F (38)
fs
where f s is the gas friction factor for a smooth wall tube (= 0.046 Re−0.2
G ).
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 34

34 Modeling of Gas Liquid Flow in Pipes

G, F and γ are given by:


ρL Dg
G= (39)
ρG UGS2
fs
 1/2
γ ReLF νL ρL
F= (40)
0.9
ReG νG ρ G
and
 
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

2.25 0.4
γ = 0.42ReLF
1.25
+ 2.8 × 10−4 ReLF (41)
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Using the information given in Eqs. (19)–(41), Eq. (18) is solved for the film
thickness δ (or ALF ).
Also note that empirical correlations for the film thickness were proposed by
Henstock and Hanratty (1976).
For concurrent upward and downward flows:
δ 6.59F
= (42)
D (1 + 1400F)1/2
for horizontal flow:
δ 6.59F
= (43)
D (1 + 850F)1/2
where F and γ are given by Eqs. (40) and (41).

2.2. Two-dimensional solution for the film


The empirical correlations for the wall shear stress, described previously are based
on the average velocity of the liquid in the film. A somewhat improved accuracy
can be obtained by solving the two-dimensional velocity profile in the film.
Figure 9 shows the geometry of an upward co-current annular flow.
For laminar film flow, a force balance on the cross hatched control volume yields
 
du L dP
τ =µ = τi − + ρg (δ − y) (44)
dy dx
The solution for the velocity profile is:
  
τi 1 dP y2
uL = y − + ρg δy − (45)
µ µ dx 2
and the average velocity in the liquid film is:
δ   2
1 1 δ dP δ
UL = u L dy = τi − + ρg (46)
δ µ 2 dx 3
0
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 35

Hydrodynamic Models Based on Flow Patterns 35


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 9. Geometry of annular flow with velocity profile in the film.

For turbulent film flow, the force balance yields


 
du L dP
τ = (µ + ε) = τi − + ρg (δ − y) (47)
dy dx
where ε is the eddy viscosity. Close to the wall, the eddy viscosity is correlated
using Deissler correlation and away from the wall, Von Karman correlation is
applied. The solution of the velocity profile is obtained numerically.
 2 
n uL y
ε = n u L y 1 − exp −
2
Deissler y + < 20, n = 0.1 (48)
νL
where ν is the kinematic viscosity,
(du L /dy)3 +
ε = k2  2 Von Karman y > 20, k = 0.36 (49)
d u L /dy
2 2


where y + = u ∗ y/νL u ∗ = τW /ρL
The gas core is treated as before, namely the calculations of the interfacial
shear stress τi and the pressure drop dP/dx are based on the average gas velocity
[Eqs. (28)–(33)].

3. Dispersed Bubble Flow


Dispersed bubble flow is the closest flow pattern to homogeneous flow. If the bubbles
move at the liquid velocity, the analysis of the homogeneous model is valid. For
high mixture flow rates this is indeed the case, the bubbles are carried by the liquid
attaining the liquid velocity. Yet, there are cases that the bubbles move at a different
velocity and there is a slip between the velocity of the bubbles and that of the liquid.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 36

36 Modeling of Gas Liquid Flow in Pipes

Typically, this happens for vertical flows where buoyancy drives the bubbles upward,
faster than the liquid velocity.
The main variables of interest in bubble flow are: pressure drop, void fraction and
void fraction distribution, bubble size and bubble size distribution, bubble velocity
and bubble velocity distribution.
The pressure drop of the mixture is
dP 4 fm ρm Um2
− = + ρm g sin β (50)
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

dx D 2
where the mixture velocity is:
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Um = ULS + UGS (51)

and the mixture density is:

ρm = (1 − α)ρL + αρG (52)

The friction factor f m , for smooth pipes is


 −n
DUm ρm
fm = CRe−n =C (53)
m
µm
The values of C and n depend on the flow (laminar or turbulent).
The void fraction is obtained using
UGS ULS
α= , (1 − α) = (54)
Ub UL
Ub is the bubble velocity. Note that for homogeneous flow Ub = UL = Um .

3.1. Bubble rise velocity


The terminal rise velocity of a single bubble in stagnant liquid U0 , depends on
the fluid properties and the bubble diameter. The dependence of the rise velocity
on the bubble radius Rb , for an air bubble in water is shown in Fig. 10. Small
bubbles are almost perfect spheres due to surface tension effect and they behave
as solid spheres (Stokes’ solution, region A–B in Fig. 10). As the bubble diameter
increases, viscosity and internal motion of the gas affect the rise velocity (region B–
C). For larger bubbles the effect of surface tension on the rise velocity is negligible
(regions C–D–E). In region A–C, the bubble rise velocity increases with the bubble
diameter, since buoyancy is proportional to the volume (∼D 3 ) and the resistance
to the upward motion is proportional to the bubble surface area (∼D 2 ); thus, as the
bubble diameter increases the effect of buoyancy versus resistance increases and so
is the bubble rise velocity. Larger bubbles (region C–E) are distorted and have a flat
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 37

Hydrodynamic Models Based on Flow Patterns 37


U0 (cm/sec)
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Equivalent radius, Rb(cm)


Figure 10. Terminal velocity of air bubbles in distilled water (Wallis, 1969).

cap shape. In this case, the rise velocity becomes almost independent of the bubble
equivalent diameter.
Harmathy (1960) proposed the following correlation for the rise velocity of a
large single bubble in stagnant liquid:

g(ρL − ρG )σ 1/4
U0 = 1.53 (55)
ρL2
Note that in this region the rise velocity is independent of the bubble diameter. This
simplifies the prediction of pressure drop and void fraction, since there is no need
to consider the bubble size.
The rise velocity of a bubble in a swarm of bubbles is lower. For this case,
Harmathy (1960) recommended to use
1/4
2 g(ρL − ρG )σ
U0 = 1.53(1 − α) (56)
ρL2
The average velocity of a swarm of bubbles rising in a flowing liquid depends on
the mixture velocity and on the cross-sectional concentration distribution

Ub = C 0 US + U0 (57)

where C0 is the Zuber–Findlay distribution parameter. The value for C0 is


approximately 1.15.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 38

38 Modeling of Gas Liquid Flow in Pipes

Thus,
1/4
g(ρL − ρG )σ
Ub = 1.15(ULS + UGS ) + 1.53(1 − α) 2
(58)
ρL2
A slightly different result is proposed by Govier and Aziz (1972).

g(ρL − ρG )σ 1/4
Ub = 1.15(ULS + UGS ) + 1.41(1 − α)3/2 (59)
ρL2
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

4. Slug Flow
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Gas liquid slug flow is an intermittent flow which exists in the whole range of pipe
inclinations and over a wide range of gas and liquid flow rates. In vertical slug
flow, most of the gas is located in large bullet shaped bubbles, which occupy most
of the pipe cross-section. These bubbles are usually called “Taylor bubbles”. The
Taylor bubbles are separated by liquid slugs containing usually small bubbles. The
liquid confined between the Taylor bubble and the pipe wall flows around the bubble
as a thin falling film. On the average vertical slug flow retains axial symmetry. In
horizontal and inclined flows, slugs of liquid which fill the whole cross-section of
the pipe are separated by a stratified zone with an elongated gas bubble in the upper
part of the pipe and a liquid film at the bottom. When the flow is calm and the liquid
slug is almost free of gas bubbles, the pattern is termed elongated bubble flow. For
high flow rates, when the liquid is aerated with gas bubbles, the flow is designated
as slug flow. In spite of the distinction between slug and elongated bubble flows, the
term slug flow is still often used for the general case of intermittent flow.
Slug flow is a highly complex type of flow with an unsteady nature, thus the
prediction of pressure drop, heat and mass-transfer for such flow is a difficult task.
Therefore, a variety of approximate methods have been developed for calculating
the slug hydrodynamic parameters. The former methods used usually correlations
based on experimental results.
Later, approximate models were proposed that are capable of simulating the
flow behavior sufficiently accurate so that the calculation of the pressure drop
as well as other flow parameters can be performed with a relatively high degree
of confidence. Such models were introduced by Dukler and Hubbard (1975) and
Nicholson et al. (1978) for horizontal flow; Fernandes et al. (1983), Sylvester (1987)
and Orell and Rembrand (1986) for vertical flow and Bonnecaze et al. (1971) for
inclined flow. Some more comprehensive approaches to slug flow modeling for the
whole range of pipe inclination were introduced by Taitel and Barnea (1990) and
Fabre and Linè (1992).
All of the afore-mentioned models deal with developed steady slug flow. There
are, however, more complex types of slugs that are of a typical transient nature.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 39

Hydrodynamic Models Based on Flow Patterns 39

It takes place in the developing entrance region and in hilly terrain pipes with
pipe inclination variation having top and bottom elbows. Note however, that even
developed slug flow is intrinsically an unsteady flow.
The purpose of modeling steady slug flow is to be fairly close to the true physical
process that is taking place, but it is also important for a model to be simple enough so
that practical solutions for the slug flow parameters can be obtained with reasonable
efforts and that the modeling could be used in routine engineering calculations.
A schematic geometry of slug flow is shown in Fig. 11. The slug unit of length
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

u , is subdivided into two main sections: the liquid slug zone of length s and the
film zone of length f . Although the liquid slug zone can be aerated by dispersed
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

bubbles, it forms a competent bridging of liquid. The liquid holdup within the
liquid slug zone is designated as Rs . Once the slug is incapable of forming a
competent bridging, the slugs are then termed proto-slugs (Andritsos and Hanratty,
1987b) and this is the beginning of transition to annular flow. The average liquid
velocity in the liquid slug body is designated as UL . The average axial velocity of
the dispersed bubble in the liquid slug is termed Ub . Note that UL and Ub are not
necessarily the same, even though for horizontal flow, both velocities are considered
equal.

Figure 11. Slug flow geometry.


July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 40

40 Modeling of Gas Liquid Flow in Pipes

The film zone consists of a liquid film and an elongated gas bubble. For
horizontal and inclined pipes, the bubble is in the upper part of the pipe. In vertical
and off vertical pipes, a complete symmetry is assumed, the elongated bubble
(Taylor bubble) is in the center of the pipe and a thin film flows around it adjacent to
the pipe wall. The translational velocity of the elongated bubble, Ut , is the velocity
at which the elongated bubble propagates downstream. If one moves at a velocity
Ut , the slug picture is seen as frozen in space. The liquid velocity in the film is
designated as Uf and the velocity of the gas in the elongated bubble UG . Note that in
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

the present approach, Rs is considered axially homogeneous while the film thickness,
hf (z) (δ(z) for the symmetrical vertical case), and its velocity, Uf (z), vary along the
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

elongated bubble.
The analysis is performed with reference to a coordinate system that is attached
to the front of the elongated bubble moving at a velocity Ut . The relative velocities
are designated as VG = Ut − UG and Vf = Ut − Uf . The coordinate z is an axial
coordinate that started at the nose of the elongated bubble directed upstream.
For the given parameters specified below (input)

• pipe diameter D
• angle of inclination β
• fluid properties ρL , ρG , µL , µG , σ
• flow rates ULS , UGS
The following slug variables are calculated (output)

• pressure drop dP/dx


• lengths: slug unit, liquid slug, film u , s , f
• slug frequency ν
• liquid holdup in the liquid slug Rs
• film thickness or film holdup hf (z), Rf (z)
• average void fraction, slug unit αu
• fluids velocities UL , Ub , Uf (z), UG (z)
• translational velocity Ut
The model detailed below is based on Taitel and Barnea (1990).

4.1. Mass balances


Mass balances presented here consider both the liquid and gas as incompressible.
For long pipelines where the density is not constant, we can still consider it as locally
constant for the purpose of calculating steady slug flow.
A liquid mass balance over a slug unit can be performed in two ways. One way
is to integrate the fluid flow rate at a fixed cross-section over the time of the passage
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 41

Hydrodynamic Models Based on Flow Patterns 41

of a slug unit. The second one is by considering the volume of the liquid in a slug
unit. Both methods obviously yield the same results.
Using the first approach for the liquid mass balance yields the following
equation:
 
tf
1
WL = ρL UL AR s ts + ρL Uf ARf dt  (60)
tu
0
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

where WL is the input liquid mass flow rate, UL is the liquid average velocity in
the liquid slug and Uf is the liquid velocity in the film. tu , ts and tf are the times for
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

the passage of the slug unit, the liquid slug and the film zone, respectively. Since
ts = s /Ut and tf = f /Ut Eq. (60) takes the form:

 f
s 1
WL = ρL UL AR s + ρL Uf ARf dz (61)
u u
0

The second way of formulating the mass balance yields the following relation,
 
 f
1
WL = ρL AR s s + ρL ARf dz  − X (62)
tu
0

The term in the square parenthesis is the mass of the liquid in a slug unit. A slug
unit is propagating in the pipe at a velocity Ut and the time for a slug unit to pass
through a fixed point in the pipe is tu = u /Ut . However, at this time, part of the
liquid in the film moves upstream (backward) relative to the gas–liquid interface
and is captured by the following slug. This amount of picked up liquid X , is given
by the expression

X = (Ut − UL )ρL AR s = (Ut − Uf )ρL ARf (63)

Using Eq. (63) for X one can show that both Eqs. (61) and (62) are indeed the same.
Equations (61) and (63) can be combined to yield

 f
f Ut
ULS = UL Rs + Ut (1 − Rs ) − αf dz (64)
u u
0

Equivalent mass balances can be performed on the gas. However, it is more


convenient to use a mass balance on one species only and a mass balance on the
mixture. A very simple continuity balance on both liquid and gas states that for
constant densities, the volumetric flow rate through any cross-section is constant.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 42

42 Modeling of Gas Liquid Flow in Pipes

Applying this balance on a cross-section in the liquid slug zone yields

Us = ULS + UGS = UL Rs + Ub αs (65)

where Us is the mixture velocity within the liquid slug and Ub is the velocity of the
bubbles in the liquid slug.

4.2. Average void fraction


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

The average void fraction of a slug unit is defined as:


 
 f
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

1 
αu = αs s + αf dz  (66)
u
0

Using the mass balance [Eq. (64)] to eliminate the integral term in Eq. (66) yields:
1 
αu = −ULS + UL Rs + Ut αs (67)
Ut
Using relation (65) the liquid flow rate ULS can be replaced by UGS to yield:
1 
αu = UGS − Ub αs + Ut αs (68)
Ut
Equations (67) and (68) present indeed a very interesting result. It shows that
the average void fraction of a slug unit depends only on the liquid and gas flow
rates (UGS , ULS ), the dispersed bubble velocity Ub , the translational velocity Ut
and the void fraction within the liquid slug αs . It is independent of the bubble shape,
the bubble length, the liquid slug length and film thickness. This is a very important
and convenient result, since it shows that the gravitational pressure drop can be
calculated, independent of the detailed slug structure. For the simple case when
the liquid slug is not aerated, Rs = 1 and Eq. (68) reduces to the simple result
αu = UGS /Ut .

4.3. Hydrodynamics of the liquid film


The length of the liquid film f , its shape hf (z), the velocity profile along the liquid
film Uf (z) and the film thickness and its velocity just before pickup h fe and Ufe , are
important parameters for calculating the pressure drop and heat and mass-transfer
in slug flow.
The shape of the liquid film is a very complex structure, especially near the tail
of the liquid slug. It is a three-dimensional problem of turbulent nature and with
free surface. Obviously an exact solution is out of question at this time. A reasonable
approximation is to use the one-dimensional approach of the “channel flow” theory.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 43

Hydrodynamic Models Based on Flow Patterns 43

This method has been used by Dukler and Hubbard (1975) and Nicholson et al.
(1978).
In order to find solutions for the film velocity Uf and the film holdup Rf as a
function of position from the rear of the slug z, we consider momentum balances
on the film zone. Referring to Fig. 11, the momentum equations for the liquid film
and the gas above it relative to a coordinate system moving with a velocity Ut are
(see Sec. 1, Chapter 6 for detailed derivation):

∂ Vf ∂P τf Sf τi Si ∂hf
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

ρL Vf =− + − + ρL g sin β − ρL g cos β (69)


∂z ∂z Af Af ∂z
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

and
∂VG ∂P τG SG τi Si ∂hf
ρG VG =− + + + ρG g sin β − ρG g cos β (70)
∂z ∂z AG AG ∂z

where Vf = Ut − Uf and VG = Ut − UG . Note that although these equations are


written for the relative velocities Vf and VG , the shear stresses are given in terms of
the real velocities as follows (see Fig. 11 for the definition of the positive direction
for the shear stresses):

τf = ff |Uf |Uf /2 (71)

τG = f G |UG |UG /2 (72)

and

τi = f i ρG |UG − Uf |(UG − Uf )/2 (73)

where ff , f G and f i are the friction factors between the liquid and the wall, the gas
and the wall and at the gas–liquid interface, respectively. Uf and UG are considered
positive in the downstream (x) direction.
For smooth pipes, the Blasius’ correlation can be used,
 −n
Dh Uf ρL
ff = Cf (74)
µL

where Dh = 4Af /Sf . A similar expression can be used for the gas, with the exception
that the gas hydraulic diameter is taken as Dh = 4AG /(SG + Si) (Taitel and Dukler,
1976). For laminar flow Cf = 16 and n = 1, while for turbulent flow Cf = 0.046
and n = 0.2.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 44

44 Modeling of Gas Liquid Flow in Pipes

For rough pipes, the roughness of the pipe should be taken into account. An
example for such an expression is the convenient explicit formula (Hall, 1957):
 1
6 3
ε 10
ff = 0.001375 1 + 2 × 104 + (75)
Dh Re

Obviously many other correlations can be used.


More problematic is the determination of the interfacial friction factor f i . In
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

the case of low liquid and gas velocities, the smooth surface friction factor can be
used. When the interface is wavy, the wavy structure determines the value of the
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

average friction factor. Unfortunately, due to the complexity of the wavy structure,
the interfacial friction factor cannot be predicted accurately and one has to use some
crude correlations and assumptions. The nature of the interface (smooth or wavy)
can be determined on the basis of flow pattern maps, considering the film zone as
stratified flow with the appropriate flow rates of liquid and gas.
For wavy stratified flow in the horizontal and inclined cases, a constant value of
f i = 0.014 was suggested (Cohen and Hanratty, 1968; Shoham and Taitel, 1984).
For the vertical case, Wallis’ correlation (1969) for co-current annular flow

δ
f i = 0.005 1 + 300 (76)
D
can be used, though the flow in the film zone is usually counter current. It may be
noted that Wallis et al. (1978; 1979) suggested modified correlations for counter
current flow. However, those correlations are applicable near the flooding point
and this is usually not the case for “normal” slug flow. As can be seen, the
information regarding the interfacial shear is very limited, primarily for inclined
pipes. Fortunately, the accuracy of the interfacial friction is generally not important
since in most cases the interfacial shear in the film zone is negligible.
Eliminating the pressure gradient from Eqs. (69) and (70) yields:
 
∂Vf ∂ VG τf Sf τG SG 1 1
ρL Vf − ρG VG = − − τi Si +
∂z ∂z Af AG Af AG
∂hf
+ (ρL − ρG )g sin β − (ρL − ρG )g cos β (77)
∂z
Using Eq. (63), the relative velocities Vf and VG are given by:
Rs
Vf ≡ (Ut − Uf ) = (Ut − UL ) (78)
Rf
Likewise
αs
VG ≡ (Ut − UG ) = (Ut − Ub ) (79)
αf
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 45

Hydrodynamic Models Based on Flow Patterns 45

Since Rf as well as αf are functions of hf (or δ), substituting these values in (77)
yields a differential equation for hf (or δ) as a function of z:
 
τf Sf τG SG 1 1
− − τi Si + + (ρL − ρG )g sin β
dhf Af AG Af AG
=
dz (Ut − UL )Rs dRf (Ut − Ub )(1 − Rs ) dRf
(ρL − ρG )g cos β − ρL Vf − ρG VG
Rf2 dhf (1 − Rf )2 dhf
(80)
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

where for the case of stratified film flow,



by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

 2
dRf 4 hf
= 1− 2 −1 (81)
dhf πD D
Equation (80) is also valid for vertical flow β = 90◦ , where hf is replaced by δ, SG
equals zero and
 
dRf 1 δ
=4 −2 2 (82)
dδ D D
The differential Eq. (80) is solved numerically for hf (z) and the corresponding
Uf (z) is found using Eq. (78). The integration is performed until the mass balance
of Eq. (64) is satisfied, yielding the length of the liquid film f , as well as the holdup
Rfe and the velocity Ufe at the end of the liquid film just before pickup.
For large z, the limiting value of h fe is the equilibrium liquid level h E , which is
obtained when dhf /dz = 0, namely the numerator of Eq. (80) equals zero.
Note that for aerated liquid slugs, the translational velocity in the elongated
bubbles usually exceeds that of the dispersed bubbles in the liquid bridge. In addition,
the liquid film seems to be essentially free of small bubbles. The physical picture
that is consistent with this description is that the dispersed bubbles in the liquid slug
coalesce at the nose of the elongated bubble, while gas bubbles are re-entrained
from the back of the bubble into the liquid slug. Thus the liquid holdup in the
front of the liquid film Rfi , equals the value of Rs and Ufi equals UL ; h s is the liquid
level corresponding to Rs . Thus, the integration of Eq. (80) starts normally with
hf = hfi = h s at z = 0 and hf decreases (dhf /dz > 0) from h s toward the limit of h E .
However, under certain conditions, dhf /dz may be positive. It occurs whenever
the critical level h c is less than h s , where h c is the level that equates the denominator
to zero. In this case, the liquid level reduces “instantaneously” to the critical level,
and the integration of hf starts with hfi = h c at z = 0 (see Fig. 12). This procedure is
similar to the analysis of liquid drainage from a reservoir to a super critical channel
flow (Henderson, 1966). We may further note that in the event that h c or h s are less
than the equilibrium level h E , then h E is reached immediately. For the vertical case,
the denominator is never zero and a critical film thickness does not exist.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 46

46 Modeling of Gas Liquid Flow in Pipes


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

Figure 12. The integration of the film level.


by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Equation (80) is the most detailed form of the one-dimensional channel flow
approach. This approach with several degrees of simplification has been used by var-
ious investigators. Dukler and Hubbard (1975) and Nicholson et al. (1978) assumed
that the pressure drop in the film zone is negligible. Under this assumption, the liquid
is treated as an uncoupled free surface channel flow and Eq. (77) takes the form
∂ Vf τf Sf ∂hf
ρL Vf = + ρL g sin β − ρL g cos β (83)
∂z Af ∂z
Equation (83) is still a differential equation that has to be integrated numerically
and the neglect of the pressure drop along the gas bubble, although usually justified,
may be incorrect for very long film zones in which the contribution of the pressure
drop in the gas zone is not negligible.
Further simplifications have been proposed in order to avoid the numerical
integration. The most common approach is to consider the liquid film as having
a constant thickness in equilibrium. This equilibrium level is indeed the solution
of h E . In the case of a vertical pipe this was, in fact, the only approach taken by
Fernandes et al. (1983), Sylvester (1987) and Orell and Rembrand (1986). In this
case, the solution for h (or δ for vertical symmetrical flow) should satisfy
 
τf Sf τG SG 1 1
− − τi Si + + (ρL − ρG ) g sin β = 0 (84)
Af AG Af AG
In the afore-mentioned works, however, the pressure drop in the gas zone was also
neglected. This neglect is usually justified only for relatively short bubbles.
In summary, several approaches with various degrees of simplicity have been
presented for the hydrodynamics of the liquid film. We focus our attention on three
cases:
Case 1: This is the most general formulation for the one-dimensional channel flow
approximation. It is given by Eq. (77) or (80).
Case 2: The liquid film is treated as a free surface channel flow (Eq. (83)).
Case 3: A uniform film is assumed along the bubble zone (Eq. (84)).
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 47

Hydrodynamic Models Based on Flow Patterns 47

4.4. Pressure drop


Since the slug is not a homogeneous structure, the local axial pressure drop is not
constant. For practical purposes, we need the average pressure drop over a slug unit,
namely Pu / u .
The pressure drop for a slug unit can be calculated using a global force balance
along a slug unit between cuts A-A and C-C (Fig. 11). The momentum fluxes in and
out are identical and the pressure drop across this control volume is:
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

 f
τs πD τs Sf + τG SG
Pu = ρu g sin β u + s + dz (85)
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

A A
0

where ρu is the average density of the slug unit:

ρu = αu ρG + (1 − αu )ρL (86)

The first term on the right-hand side (RHS) of Eq. (85) is the gravitational
contribution to the pressure drop, whereas the second and third terms are the
frictional terms in the slug and in the film zones.
A second method, which is frequently used for calculating the pressure drop
is to neglect the pressure drop in the film region and to calculate the pressure drop
only for the liquid slug zone. In the slug zone, a control volume between the plane
cuts A-A and B-B is used. The resulting pressure drop along a slug unit, Pu , in
this case is
τs πD
Pu = ρs g sin β s + s + Pmix (87)
A
where ρs is the average density of the liquid slug body, namely:

ρ s = αs ρ G + R s ρ L (88)

The first term in the RHS of Eq. (87) is the gravitational term of the liquid slug;
the second term is the pressure loss due to friction and the third term, Pmix ,
is the pressure loss in the near wake region behind the long bubble. Dukler and
Hubbard (1975), Nicholson et al. (1978) and Stanislav et al. (1986) proposed that
this pressure drop is associated only with the acceleration of the liquid in the film
to the liquid velocity within the liquid slug, namely

Pmix = Pacc = ρL Rs A(Ut − UL )(UL − Ufe ) (89)

However, a careful momentum balance between cuts A-A and C-C indicates that
the contribution to Pmix , is not only due to the acceleration pressure drop, but also
due to the change in the liquid level between the film zone and the liquid slug zone
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 48

48 Modeling of Gas Liquid Flow in Pipes

(Taitel and Barnea, 1990), namely:


h fe hfi
APmix = ρL g cos β (h fe − y)b dy−ρL g cos β (hfi − y)b dy
0 0
+ ρL Rs A(Ut − UL )(Ufi − Ufe ) (90)
where y and b are defined in Fig. 4, Chapter 6.
As discussed in the previous section, hfi is usually h s and Ufi is UL , in which
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

case the last term on the RHS of Eq. (90) is equal to the acceleration pressure drop
as given by Eq. (89). However, when h c < h s , hfi equals h c , Ufi = Uc and the
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

acceleration term in Eq. (90) is different from that in Eq. (89). The integration in
Eq. (90) can be carried out and written explicitly.
h fe
(h fe − y)b dy
0


3  2 3  
D  h fe D3 h fe
= 1− 2 −1 + 2 −1
12 D 8 D
  
      2 
π h fe h fe h fe
×  + sin−1 2 −1 + 2 −1  1− 2 −1  (91)
2 D D D

Inspection of Eq. (90) shows that Pmix is always less than Pacc . The use of
Pmix = Pacc may cause a minor error for small diameter pipes but it can lead to
a serious error when the pipes are of a large diameter.
Two methods have been presented for the pressure drop calculation. In the first
method, a global force balance is used (85), while the second method is based on the
momentum balance only on the liquid slug zone, neglecting the pressure drop in the
film zone (87). It will now be shown that the two methods are identical, provided that
in the first method we also assume that the pressure along the film zone is essentially
constant.
The integrated form of the momentum balance given by Eq. (77) is:
U
t −Ufe
∂(Ut − Uf )
ρL Rf (Ut − Uf ) dz
∂z
Ut −Ufi
 f  f h fe
τf Sf ∂hf
= dz + ρL g sin β Rf dz − ρL g cos β Rf dz (92)
A ∂z
0 0 hfi

The left hand side (LHS) of Eq. (92) is exactly the acceleration term Pacc .
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 49

Hydrodynamic Models Based on Flow Patterns 49

Integrating by parts, one can show that:

hf hf
Af dhf = (hf − y) b dy (93)
0 0

Equation (92) then takes the form:

ρL Rs A(Ut − UL )(Ufi − Ufe )


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

 f  f h fe
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

= τf Sf dz + ρL g sin β Af dz − ρL g cos β (h fe − y) b dy
0 0 0

hfi
+ ρL g cos β (hfi − y) b dy (94)
0

Substituting Eq. (94) into Eq. (90) yields another expression for Pmix ,

 f  f
APmix = ρL g sin β Af dz + τf Sf dz (95)
0 0

By substituting Pmix of Eq. (95) into Eq. (87), one can see that it is identical to
Eq. (85) (for the case where pressure drop in the film zone is zero). Namely, the two
methods for the pressure drop calculation: (i) A global momentum balance on the
whole slug unit, Eq. (85) and (ii) a momentum balance over the liquid slug only,
Eq. (87), are identical.
A simplified approach that considers a uniform film thickness is frequently used
(Fernandes et al. 1983 and Sylvester, 1987), the calculations of the pressure drop
via Eqs. (85) and (87) are not consistent in this case and yield different results.
Equation (85) with a constant film thickness and negligible pressure drop in the gas
zone yields only the first two terms of Eq. (87) without the acceleration term. As
pointed out by Barnea (1990) and demonstrated by Eq. (95), one has to consider a
curved nose in order to obtain the same results for the pressure drop by using either
Eq. (85) or (87).
Barnea (1990) compared the results of the pressure drop for the constant film
thickness model with the more exact solutions using a curved nose calculation. She
showed (see Fig. 13) that for the vertical case, when a cylindrical bubble with a flat
nose is assumed, the results of the pressure drop without the acceleration term in
Eq. (87) is usually closer to the more exact solution than when the acceleration
term is used. Note that in the work of Fernandes et al. (1983) and Sylvester
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 50

50 Modeling of Gas Liquid Flow in Pipes


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 13. Comparison of pressure drop calculations with experimental results.


Vertical upward flow, air–water, 0.1 MPa, 25◦ C, 2.8 cm diameter; Mode 2: exact, Mode 3: “flat nose”,
with acceleration (Pmix ), Mode 4: “flat nose”, without acceleration (Pmix ) (Barnea, 1990).

(1987) the acceleration term was considered, although the constant film thickness
approximation was used.

4.5. Auxiliary relations


The formulations provided so far are not sufficient yet to obtain a solution. In order
to proceed, we should consider the following additional auxiliary variables: (1)
the translational velocity Ut , (2) the dispersed bubbles velocity within the liquid
slug Ub , (3) the liquid holdup in the liquid slug zone Rs , and (4) the liquid slug
length s , or the slug frequency ν. These variables are provided by relations based
on experimental results and modeling related to each variable. These relations are
updated continuously in the literature and they can be used in the solution procedure.

4.5.1. The translational velocity


The translational velocity Ut , is the velocity of the elongated bubble front interface.
For aerated liquid slugs, it is larger than the gas velocity of the elongated bubble
UG . The prediction of the translational velocity is not an easy task and in fact it is
still the subject of current research.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 51

Hydrodynamic Models Based on Flow Patterns 51

It is generally assumed that the translational velocity of a single elongated bubble


in a pipe is a superposition of the bubble velocity in a stagnant liquid i.e. the drift
velocity Ud , and the contribution due to the mean liquid velocity UL (Nicklin et al.,
1962).

Ut = CUL + Ud (96)

where C is a coefficient that depend on the liquid velocity profile. It is also termed
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

the distribution parameter, owing to its similarity to the coefficient proposed by


Zuber and Findlay (1965). For continuous slug flow, Nicklin et al. (1962) proposed
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

to use the above correlation while substituting the mean liquid velocity UL by the
mixture velocity Um (= Us ).

The distribution parameter C


The value of the constant C is based on the assumption that the propagation
velocity of elongated bubbles follows the maximum local liquid velocity Umax ,
in the liquid in front of the nose tip, namely C = Umax /UL (Nicklin et al., 1962;
Grace and Clift, 1979; Bendiksen, 1984; Shemer and Barnea, 1987). The parameter
C equals approximately 1.2 for fully-developed turbulent flow and 2 for laminar
flow. This assumption was confirmed by Polonsky et al. (1999b) by direct particle
image velocimetry (PIV) measurements of the cross-sectional velocity profiles in
front of the Taylor bubble, together with independent measurements of the Taylor
bubble translational velocity in stagnant, vertical upward and downward liquid flow
as shown in Fig. 14. For upward liquid flow its maximum is at the pipe center.
However for downward (negative) flow the maximum is at the pipe wall where the
velocity is zero. As a result, the nose of the Taylor bubble tends to move toward the
pipe wall.

Figure 14. The relation between C and the velocity profiles ahead of the bubble (Polonsky et al.,
1999b).
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 52

52 Modeling of Gas Liquid Flow in Pipes


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 15. Effect of the mixture flow velocity on the coefficient C.


Source:  Fréchou (1986);  Mao and Dukler (1991); — empirical fit (Fréchou, 1986).

Figure 15 presents experimental results, reported by Fabre and Linè (1992), for
the coefficient C as a function of the Reynolds number.
They suggested the following empirical fit.

2.27 1.2
C= + (97)
1 + (Re/ReC ) 2 1 + (ReC /Re)2

where ReC is a critical Reynolds number ≈ 1000.


The dependence of the factor C (Eq. (96)) on the inclination angle was correlated
by Bendiksen (1984) for a pipe diameter of 2.42 cm.

C(β) = C(0◦ ) + [C(90◦ ) − C(0◦ )] sin2 β (98)

For the inertial regime, Collins et al. (1978) solved analytically the problem of a
bubble rising in an up-flowing liquid. The assumption was made that the liquid
around the bubble is inviscid and that the role of viscosity is limited to creating
the velocity profile far ahead of the bubble. Laminar and turbulent flows were
considered. In the case of laminar flow, they reported a slight higher value for
the coefficient C compared to Nicklin’s suggestion, C = 2. Later Bendiksen (1985)
showed that this over-prediction is due to the neglect of the surface tension. In the
case of turbulent flow, Collins’ coefficient C agrees approximately with Nicklin’s
proposal (C = 1.2).

The drift velocity


A large part of the research efforts on the propagation velocity of elongated bubbles
has been focused on the buoyancy drift velocity Ud , namely the velocity of the
bubbles in stagnant liquid.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 53

Hydrodynamic Models Based on Flow Patterns 53

Usually, the asymptotic cases are considered: inertial flow, viscous flow and
capillary flow. The ranges of these regimes were determined experimentally by
White and Beardmore (1962). Three dimensionless numbers are of importance in
the case where density and viscosity of the gas are small compared to those of
the liquid: Froude number, which represents the ratio of inertial to gravitational

forces, Fr = Ud / g D; Eötvös number, representing the ratio of gravitational
forces to surface tension, Eo = ρL g D 2 /σ and the property group, which contains
the liquid properties, Y = gµ4L /ρL σ 3 . White and Beardmore (1962) found that
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

inertial effects are negligible


 for Fr < 0.003, interfacial effects for Eo > 70
and viscous effects for Eo3 /Y < 3 × 105 . Wallis (1969) plotted similar ranges
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

for somewhat different dimensionless groups. Wallis (1969) suggests to correlate


Fr (defined as Fr = Ud /[(ρL − ρG )g D/ρL )]1/2 ) with respect to the Eötvös
number, Eo = g D 2 (ρL − ρG )/σ and the inverse viscose number, Nf = [D 3 g
(ρL − ρG )ρL ]1/2 /µL .
 0.01Nf
 (3.37−Eo)

Fr = 0.345 1 − e− 0.345 1 − e m (99)

where m = 10 for Nf > 250, m = 69 Nf−0.35 for 18 < Nf < 250 and m = 25 for
Nf < 18.
The drift velocity depends on the three-dimensional flow at the front of the
bubble. In the case of inertial flow (where both interfacial and viscous effects are
negligible), the fluid can be assumed inviscid.
Davies and Taylor (1949) considered a potential flow around a spherical nose
of a rising bubble in stagnant liquid

2 9
U 2 = U∞ sin2 ϕ (100)
4
where U is the tangential liquid velocity on the surface of the bubble, ϕ is a polar
coordinate and U∞ is the free stream velocity, which is equal to the drift velocity
Ud for a rising bubble in a stagnant liquid.
Application of the Bernoulli’s equation between the bubble nose and the base of
the cap and considering the pressure within the bubble as constant yields gz = U 2 /2.
Substituting this relation in (100) yields for small ϕ:

4 
Ud = g D b = 0.471 g D b (101)
18
In Eq. (101), Db is the bubble diameter. Davis and Taylor (1949) modified this
equation to take into account the bubble rise velocity in a confined pipe. They used a
series of expansion techniques and obtained a constant of 0.328 instead of 0.471 and
the bubble diameter Db was replaced by the pipe diameter D. Dumitrescu (1943)
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 54

54 Modeling of Gas Liquid Flow in Pipes

performed somewhat more accurate calculations and obtained the constant of 0.35
which thereafter was accepted as the best value that also agrees very well with
experimental observations (Nicklin et al., 1962; Goldsmith and Mason, 1962; Clift
et al., 1978).
A comprehensive review on the rise velocity of long gas bubbles in round pipes
is reported by Viana et al. (2003). Based on their own experiments and data from
the literature, they provided a “Universal correlation” for the Froude number as a
function of the buoyancy Reynolds number (equal to Nf ) and the Eötvös number.
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

One may conclude that the drift velocity is controlled via a relation among 3
dimensionless parameters (for a given inclination). All combinations of the afore-
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

mentioned three parameters are identical in principal.


For the horizontal case, the elongated bubbles are not symmetric. A theoretical
analysis for calculating the drift velocity in horizontal pipes was presented by
Benjamin (1968). Consistent with the approach taken for the vertical case, in the
horizontal case also, the inviscid theory is applied near the bubble nose region. The
drift velocity in horizontal slug flow is the same as the velocity of the penetration
of a bubble when liquid is emptied from a horizontal tube (Benjamin, 1968).
For a coordinate system attached to the gas liquid interface (Fig. 16), the
following relations are obtained
Continuity:

AV1 = A2 V2 (102)

where A2 is given by
 
1
A2 = π − γ + sin 2γ R 2 (103)
2
Bernoulli’s theorem is applied between point (1) and the stagnation point (0). Note
that the pressure at the stagnation point, which is the same as the pressure in the gas
bubble is taken as the reference pressure.
1
P1 = − ρV12 (104)
2

Figure 16. Propagation of a gas front in a draining horizontal pipe.


July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 55

Hydrodynamic Models Based on Flow Patterns 55

Bernoulli’s theorem between point (0) and point (2) along the free surface yields:
V22 = 2g R(1 − cos γ) (105)
Finally a momentum balance yields:
h 2
(P1 + ρg R)πR − 2
ρg(h − y)b dy = ρV2 A2 (V2 − V1 ) (106)
0
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

where h 2 is the film thickness at point 2. The integral term in Eq. (106) can be solved
explicitly, namely
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

h 2  
2
ρg(h − y)b dy = ρg R A2 cos γ + R 2 sin3 γ (107)
3
0

As shown in Benjamin (1968), Eqs. (102)–(106) are solved for the liquid level h 2 ,
(equivalent to A2 , or γ) and the liquid velocities V1 and V2 . The results are
h2 
= 0.563 and V1 = Ud = 0.542 g D (108)
D
The result of Eq. (108) is supported experimentally (for elongated bubbles with a
negligible surface tension) by Zukoski (1966) and Bendiksen (1984). It is interesting
to observe that the drift velocity in the horizontal case is larger than the drift velocity
in the vertical case.
For the inclined case, one relies primarily on experimental data. The inclined
case, as well as the vertical and the horizontal cases, were studied by Zukoski (1966),
Singh and Griffith (1970), Bonnecaze et al. (1971), Bendiksen (1984) and Hasan
and Kabir (1986). All report a “peculiar” behavior, that the drift velocity increases
as the angle of inclination is declined from the vertical position. The drift velocity
then decreases again toward the horizontal position such that the maximum drift
velocity occurs at an intermediate angle of inclination around 40◦ to 60◦ from the
horizontal. Bonnecaze et al. (1971) were the first to give a qualitative explanation for
this peculiar behavior, arguing that the gravitational potential that drives the liquid
velocity along the curved surface at the bubble nose increases and then decreases
as the angle of inclination changes from the vertical position toward the horizontal
position.
Figure 17 shows the experimental data reported by Zukoski (1966). It shows
that the effect of surface tension can be indeed substantial, particularly for small
diameter pipes. For a small surface tension parameter  = 0.001, the results
for the vertical case and the horizontal case are very close to the potential flow

theory, namely Ud / gD = 0.35 and 0.54, respectively. The drift velocity, however,
decreases considerably with an increase in the surface tension parameter (decreasing
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 56

56 Modeling of Gas Liquid Flow in Pipes


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 17. Bubble drift velocity versus surface tension parameter for β = 0◦ , 45◦ , 90◦ (Zukoski,
1966).

Figure 18. Bubble drift velocity versus surface tension parameter for various Reynolds numbers.
Source: Symbols from Barr (1926), Dumitrescu (1943) and Goldsmith and Mason (1962); (Zukoski,
1966).

the pipe diameter) and eventually reaches a zero velocity when  is of the order
of unity.
The drift velocity is expected to depend also on the liquid viscosity, or the
bubble Reynolds number. However, Zukoski (1966) shows that the dependence
of the drift velocity on viscosity is negligible for Reynolds number Re > 300
(Re = Ud ρ D/µL ). This is clearly demonstrated by Fig. 18.
Bendiksen (1984) studied the propagation velocities of single elongated bubbles
in flowing liquids at different inclination angles. The obtained velocities were plotted
against the liquid velocity for each inclination angle. The drift velocities were then
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 57

Hydrodynamic Models Based on Flow Patterns 57

determined by extrapolation of the data to zero liquid velocity. The agreement


between these drift velocities and those measured directly by Zukoski (1966) in
stagnant liquid at the appropriate surface tension parameter was within 1%. The drift
velocity for inclined flows has been correlated by Bendisen (1984) as a weighted
superposition of the drift velocity in vertical and horizontal flow.

Ud (, β) = Udh (, αB ) cos β + Udv (, αB ) sin β (109)

where αB is the void fraction in the film region of the elongated bubble. The values
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

of the drift velocity for horizontal flow Udh and the drift velocity for vertical flow
Udv in Eq. (109) can be determined accouring to Zukoski (1966) (Fig. 17). Note that
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

√ √
for negligible surface tension Udh = 0.54 g D and Udv = 0.35 g D.
Hasan and Kabir (1986) proposed the relation:
 
Ud = 0.35 g D sin β(1 + cos β)1.2 (110)

which they claim to well correlate experimental data in the range 90◦ > β > 30◦ .
Figure 19 shows the results of Bendiksen (1984) and Zukoski (1966) for the

change of the dimensionless drift velocity Ud / g D with the angle of inclination.
The upper curve represents the case where surface tension is negligible and
thus the results for the horizontal and the vertical limits follow very closely the

Figure 19. Dimensionless bubble velocity in stagnant liquid versus inclination angle for different
surface tension parameters (Bendiksen, 1984).

Notes: (—) Eq. (109); () = 0.064; ( )  = 0.042; () = 0.01 (Bendiksen, 1984); (×) =
0.001; (+)  = 0.01, 0.042, 0.064 (Zukoski, 1966).
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 58

58 Modeling of Gas Liquid Flow in Pipes

afore-mentioned theoretical results, that were based on potential flow; namely,


√ √
Ud / g D = 0.35 for the vertical case (β = 90◦ ) and Ud / g D = 0.54 for the
horizontal case. The surface tension effect is given in terms of the surface tension
parameter  = 4σ/g(ρL − ρG )D 2 .

Effect of gas compressibility


Polonsky et al. (1999a) investigated the effect of gas compressibility on the motion
of Taylor bubbles by processing consecutive series of digitized video images.
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

Measurements of both bubble nose and bubble bottom velocities were performed
along the central section of the pipe. The results are shown in Fig. 20 for the case
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

of stagnant liquid.
The nose velocity is notably higher than that of the bottom; the difference
between the two increases with the bubble length. Errors in the determination of the
corresponding velocities from the recorded images are denoted by error bars. Note
that due to vigorous oscillations of the bubble bottom, errors in the determination
of the bubble bottom velocity are substantially larger than those of the bubble head.
The observed variation of the velocities of the bubble nose and bottom with the
bubble length can be explained on the basis of the following simple considerations.
Consider a Taylor bubble of mass m G , volume VG and density ρG , rising in a
vertical pipe open from above. As a result of expansion of the gas in the bubble due
to hydrostatic pressure gradient, an additional liquid velocity u L is induced at the
bubble front, which is related to the rate of bubble expansion:
   
dV G d mG d 1 VG dρG
uL A = = = mG =− (111)
dt dt ρG dt ρG ρG dt

Figure 20. Bubble nose and bottom velocities for stagnant liquid; points denote experimental data,
solid lines-model (Polonsky et al., 1999a).
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 59

Hydrodynamic Models Based on Flow Patterns 59

where
L
VG = πr 2 (z)dz (112)
0

r(z) is the local radius of the bubble, L is the bubble length and z is the axial
coordinate measured from the bubble nose. Equation (96) for the velocity of the
bubble nose, UN is thus replaced by
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

UN = C(UL + u L ) + Ud (113)
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

According to ideal gas law,


P Pa + ρL gh
ρG = = (114)
RG T RG T
dρG ρL g dh ρL g
= =− UN (115)
dt RG T dt RG T
where Pa is the atmospheric pressure at the pipe outlet, h is the hydrostatic head in
the pipe above the bubble and RG is the air gas constant. Inserting Eqs. (111), (112)
and (115) into Eq. (113) yields
L
r 2 (z)dz
0
u L = UN   (116)
R ρPLag + h
2

and
 L
−1
 r (z)dz  2
 0 
UN (h) = (CUL + Ud ) 1 − C   (117)
 R 2 Pa
+h  ρL g

where R is the pipe radius. The velocity of the bubble bottom is


dL
UB = UN − (118)
dt
The rate of length change can be found from the rate of volume change
dV G dL 2
= πr (L) (119)
dt dt
On the other hand,
dV G d V G dh dV G
= = −UN (120)
dt dh dt dh
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 60

60 Modeling of Gas Liquid Flow in Pipes

where
 
dV G d m G RG T VG
= =− (121)
dt dh Pa + ρL gh Pa
ρL g
+h

Combining Eqs. (118)–(120) yields


L
UNr 2 (z)dz
dL 0
=  
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

(122)
dt r 2 (L) ρPLag + h
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

and
 L

 r (z)dz 
2
dL  0 
UB = UN − = UN 1 −   (123)
dt  r 2 (L) ρPLag + h 

The shape of the bubble r(z) is calculated using (83). The values of UN and UB ,
calculated according to Eqs. (117) and (123) are also presented in Fig. 20. The model
and the experimental results compare favorably.
Polonsky et al. (1999b) measured the induced velocity, u L profile in front of the
Taylor bubble in stagnant liquid. The measured induced velocities are quite small
(1 to 6 mm/s at the center line) and they increase with bubble length (L/D = 2.5
to 14). The contribution of the velocity induced by the bubble expansion can thus
be ignored for nonzero liquid flow rate. The value of C or Umax /UL for the induced
velocity is also plotted on Fig. 14.

Elongated bubbles velocity in continuous slug flow


van Hout et al. (2002b) measured the translational velocities of elongated bubbles
in continuous slug flow for various flow rates, pipe inclinations and pipe diam-
eters. Measurements were carried out by cross-correlating the output signals of
consecutive optical fiber probes and by image processing techniques. In addition,
the velocities of single elongated bubbles in a stagnant and in flowing liquid were
measured by the same techniques. For all cases the measured velocities were
compared to the Nicklin’s Eq. (96) with the appropriate correlations for C and Ud
(Eq. (109)). The measured velocities of single elongated bubbles and for continuous
slug flow in the small diameter pipe (D = 0.024 m) were in all cases predicted
quite well by the afore-mentioned correlations. For continuous slug flow in the
larger diameter pipe (D = 0.054), the agreement holds only for slightly inclined
pipes (β < 20◦ ), while for larger values of β, a substantial discrepancy (up to 50%)
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 61

Hydrodynamic Models Based on Flow Patterns 61


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 21. Translational velocities of elongated bubbles in continuous slug flow as a function of
inclination angle D = 0.054 m (van Hout et al., 2002b). (a) ULS = 0.01 m/s; UGS = 0.41 m/s.
(b) ULS = 0.09 m/s; UGS = 0.34 m/s. (c) ULS = 0.21 m/s; UGS = 0.34 m/s. (d) ULS = 0.25 m/s;
UGS = 0.41 m/s. (×) Optical fiber probes; (—) (Eq. (96)); () image processing; () Ut,eff
(Eq. (126)).

is observed between the correlation results and the measured values, as shown in
Fig. 21. This discrepancy is ascribed to the influence of the dispersed bubbles in the
liquid slug region that exist in the larger diameter pipe and not in the smaller one.
van Hout et al. (2002b) found that the dependence of the measured translational
velocity of the elongated bubbles in continuous slug flow on the mixture velocity
is indeed linear as proposed by Nicklin. The coefficient C was found to be
about 1.2 for all inclinations. The drift velocity at each inclination angle was
extracted from extrapolating this dependence to Um → 0. The fact that the
estimated value of the parameter C in continuous slug flow is close to 1.2, suggests
that the observed difference between the measured elongated bubble velocity in
continuous slug flow and those predicted by the Nicklin’s correlation as observed
in Fig. 21, can only be ascribed to the use of inappropriate values for the drift
velocities. In Fig. 22, the drift velocities obtained by extrapolation of Ut versus
Um in continuous slug flow are compared to the measured drift velocities of single
bubbles. The extrapolated drift velocities are significantly higher than those of single
bubbles.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 62

62 Modeling of Gas Liquid Flow in Pipes


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 22. Froude number based on the drift velocity as a function of inclination angle. D = 0.054 m;
 = 0.010. ( ) Zukoski (1966); (—) Bendiksen (1984); (- - -) Alves et al. (1993). Single bubble data:
() image processing; (×) optical probes. Continuous slug flow: (◦) data-extrapolation, bars represent
STD. error; ( ) model range of Ud,eff , Eq. (125).

As aforementioned, the observed difference between the drift velocity of single


elongated bubbles in stagnant liquid and those in continuous slug flow (Fig. 22) can
be attributed to the presence of dispersed bubbles in the liquid slug region. When the
dispersed bubbles in continuous slug flow propagate at a velocity that is lower than
the translational velocity of the elongated bubbles, they coalesce with the elongated
bubble at its nose. At the same time dispersed bubbles are created at the elongated
bubble tail. The coalescence of the dispersed bubbles at the elongated bubble nose
results in an effectively higher translational velocity of the elongated bubble.
For simplicity, consider a flat-nosed, cylindrical elongated bubble, propagating
in stagnant liquid at a drift velocity Ud as shown in Fig. 23. In front of its nose, in
the liquid slug region with a mean void fraction of αs , dispersed bubbles move at
a drift velocity U0 . The increase of the elongated bubble drift velocity, due to the
dispersed bubble coalescence at the bubble front, can be estimated as follows:
In the absence of dispersed bubbles, the displacement of the elongated bubble
nose position during a time interval t would be x1 = Ud t. However, in the
presence of dispersed bubbles and for Ud > U0 , dispersed bubbles are absorbed
in the elongated bubble during the time interval t and the elongated bubble front
is displaced by an additional increment x2 . The liquid film region around the
elongated bubble is assumed to be free of dispersed bubbles. A simple mass balance
at the elongated bubble front yields:
αs x1 − αs U0 t
x2 = (124)
αB − αs
where αB is the average void fraction in the elongated bubble region.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 63

Hydrodynamic Models Based on Flow Patterns 63


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 23. A sketch for the model.

The effective drift velocity Ud,eff = (x1 + x2 )/t, is given by

(Ud − U0 )αs
Ud,eff = Ud + for Ud > U0 (125)
αB − αs

where the last term is the addition to the drift velocity of a single bubble due to
the absorption of the dispersed bubbles at the elongated bubble nose in continuous
slug flow. Note that this term vanishes, when αs = 0 (no dispersed bubbles in the
liquid slug) or when the elongated bubble drift velocity equals the drift velocity of
the dispersed bubbles, Ud = U0 . When Ud < U0 , dispersed bubbles are present
only in the near wake region and are absent in the front of the trailing bubble.
Hence, the void fraction in the liquid slug region, αs , will eventually become
negligible.
The effective translational velocity of elongated bubbles in continuous slug flow
is thus given by

(Ud − U0 )αs
Ut,eff = Ut,N + for Ud > U0 (126)
αB − αs

where Ut,N is the predicted translational velocity of elongated bubbles by the


Nicklin’s correlation (96).
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 64

64 Modeling of Gas Liquid Flow in Pipes

The average void fraction in the liquid slug αs , the drift velocity of a single
elongated bubble Ud , the drift velocity of a dispersed bubble U0 and the void fraction
in the elongated bubble zone αB can be determined either experimentally or by
available correlations. For more details, see van Hout et al. (2002b).
The effective translational velocity, Ut,eff , can now be calculated from Eq. (126)
for different inclination angles, flow conditions and pipe diameters. A comparison
between the measured translational velocity Ut and the predicted translational
velocity in continuous slug flow Ut,eff , is also shown in Fig. 21.
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

Note that for the smaller pipe diameter (D = 0.024 m), no significant difference
between the translational velocities of single elongated bubbles and those in
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

continuous slug flow was observed for all pipe inclinations (van Hout et al.,
2002b), since in this case Ud < U0 and the liquid slug is almost free of dispersed
bubbles.

4.5.2. Bubbles velocities in the liquid slug


The velocity of bubbles in the liquid slug Ub , is formulated similarly to the
translational velocity of elongated bubbles, namely as a linear combination of the
bubble drift velocity U0 and the mixture velocity in the liquid slug zone Us ,

Ub = BUs + U0 (127)

B is the distribution parameter (Zuber and Findlay, 1965).


The mixture velocity in the aerated liquid slugs is usually quite high and
the bubbles within the liquid slug behave as dispersed bubbles. In this case
U0
Us , B ≈ 1 and the bubble velocity equals approximately to the mixture
velocity.
For vertical and off-vertical flows and relatively large diameter pipes, discrete
bubbles also appear at low mixture velocities in the liquid slug. In this case the
bubbles within the liquid slug are not very small (compared to the dispersed bubbles)
and the free rise velocity proposed by Harmathy (1960), multiplied by sin β, is
recommended:
1
σ g(ρL − ρG ) 4
U0 = 1.53 sin β (128)
ρL2
The value of B depends on the concentration distribution of the bubbles in
the liquid slug as demonstrated by the method of Zuber and Findlay (1965).
Wallis (1969) points out that B for vertical dispersed flow usually lies between
1.0 and 1.5 with a most probable value of about 1.2. However, owing to the lack of
supporting evidence we would recommend the use of B = 1.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 65

Hydrodynamic Models Based on Flow Patterns 65

4.5.3. Void fraction in the liquid slug


The liquid slug between consecutive elongated bubbles may be aerated with small
gas bubbles. Data on void fraction within the liquid slug is usually needed for closure
of slug flow models.
A widely used empirical correlation for estimating the void fraction in the liquid
slug as a function of the mixture velocity Us , for horizontal flows was presented by
Gregory et al. (1978). They used capacitance type liquid volume fraction sensors
for air–oil slug flow in 2.58 and 5.12 cm inside diameter (I.D.) pipes and found a
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

modest diameter effect:


by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

1
Rs =  (129)
Us 1.39
1+ 8.66

Note that in their correlation Us has units of m/s.


Haywood and Richardson (1979) used γ -ray absorption method in order to
determine the average holdup within the liquid slug for an air–water system in a
4.2 cm horizontal pipeline. Their results are similar to the correlation presented by
Gregory et al. (1978).
A model to predict the average void fraction in the liquid slug was presented by
Barnea and Brauner (1985). The model is based on the assumption that the liquid
slug (at any gas and liquid flow rates) is assumed to accommodate the same gas void
fraction as bubble flow with the same mixture velocity on the transition boundary
between bubble and slug flow. Obviously, predictions based on this method rely on
correctly determined transition boundary between slug and bubble flows. Detailed
description will follow.
Barnea and Shemer (1989) used a conductance probe to detect the passage of
the gas–liquid interface at the centerline of a vertical tube. The recorded information
was further processed to obtain the void fraction at various axial locations along the
liquid slug. Andreussi and Bendiksen (1989) used a conductance probe technique to
measure the void fraction in the liquid slug in horizontal and near horizontal flows.
They compared the data to a semi-empirical correlation that predicts reasonably well,
the observed effects of diameter, inclination and physical properties. Void fraction
distribution in the liquid slug in vertical upward flow was studied experimentally
by van Hout et al. (1992). The distribution of the phases was investigated by
simultaneous measurements with two optical fiber probes, such that the distribution
was obtained over a dense grid in both the axial and radial directions. Nädler and
Mewes (1995) experimentally studied the effect of liquid viscosity on the phase
distribution in air–liquid slug flow in a horizontal pipe. The liquid holdup was
measured by a multi-beam X-ray densitometer. Reinecke et al. (1998) used a new
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 66

66 Modeling of Gas Liquid Flow in Pipes

tomographic measurement technique for the spatial imaging of the phase distribution
to measure the void fraction in horizontal slug flow.
Gomez et al. (2000) developed a correlation to predict the void fraction in the
slug body for inclination angles ranging from horizontal to vertical. The correlation
was based on 283 data points collected from six different slug flow studies that
covered a wide range of pipe diameters, fluid properties and inclination angles.
An empirical equation for estimating liquid slug holdup in horizontal and slightly
inclined two-phase flow was developed by Abdul-Majeed (2000) as a function of
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

mixture velocity, liquid viscosity and inclination angle. Cook and Behnia (2000)
performed an experimental and numerical study of the motion of bubbles in inclined
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

intermittent gas–liquid flow and reported, inter alia, on the liquid slug void fraction
at 10◦ upward inclination. Mori and Miwa (2002) measured the distribution of gas
and liquid in a two-phase vertical slug flow. Empirical correlations were proposed
for the void fraction in the wake zone, and the mean void fraction in the liquid slug.
Later, Zhang et al. (2003) developed a theoretical unified mechanistic model for the
holdup in the liquid slug, based on a balance between the turbulent kinetic energy
of the liquid phase and the surface free energy of dispersed spherical gas bubbles.
Fossa et al. (2003) used ring impedance probes in two-phase horizontal intermittent
flow to measure the instantaneous cross-sectional averaged void fraction.
Gas entrainment at the rear of a Taylor bubble that was held stationary in a
vertical tube was investigated by Delfos (1996). A model for gas entrainment was
also proposed. An experimental study was carried out in sequel by Kockx et al.
(2005) with the purpose to validate the model. Zheng and Che (2006) applied
a number of different experimental techniques to study various hydrodynamic
parameters of vertical slug flow. In particular, they presented results on the radial
and axial distributions of the void fraction in the liquid slug. Perez (2007) performed
an extensive work in investigating the behavior of air–water flow in inclined pipes
between 20◦ downwards and vertical upwards. Time series of liquid holdup (using
capacitance probes) and pressure drop (using differential pressure transducer) were
measured. In addition, a high speed video system was used to obtain image sequence
of the flow. Guet et al. (2006) developed a physically based model for computing the
mean void fraction and the liquid slug void fraction in vertical upward gas–liquid
intermittent flow. A new model for the rate of gas entrained from the Taylor bubble
to the liquid slug was formulated. It employs the framework of the Taylor bubble
wake model suggested by Brauner and Ullmann (2004).
Al-Safran (2009) developed a correlation to predict slug liquid holdup in
horizontal pipes, based on a large dataset with wide ranges of operational,
geometrical and physical property parameters. Kaji et al. (2009) reported on the
average void fraction in the liquid slug in vertical upward flow derived from wire-
mesh sensor data. Hydrodynamic characteristics like the velocity field, the volume
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 67

Hydrodynamic Models Based on Flow Patterns 67

fraction distribution of the dispersed small bubbles, the wall shear stress and the
mass transfer coefficient were investigated by Yan and Che (2011) using numerical
simulations in gas–liquid upward slug flow in a vertical pipe.
Barnea et al. (2013) performed a detailed experimental study on three-
dimensional voidage distribution in the liquid slug for various liquid and gas flow
rates and for upward pipe inclinations ranging from 2◦ to 90◦ . The measurements
were carried out using a wire-mesh sensor. About 1000–1500 liquid slugs constituted
the ensemble under investigation for each set of experimental parameters. The cross-
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

sectional average void fractions measured in the near wake region close to the rear
of the Taylor bubble have high absolute values due to the high turbulence level in
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

this region. It then gradually decreases to a nearly constant value in the far wake.
Higher inclination angles yield higher void fractions both in the wake region and
in the far wake. The near wake region that manifests itself in a sharp decrease of
the void fraction extends to distances ranging from about one pipe diameter for low
inclinations and increasing to about 2–3D for higher pipe inclination angles. For
each pipe inclination, the values of the void fraction at each location relative to the
Taylor bubble increase with the mixture velocity. The void fraction at the bottom
of the pipe in inclined flow, or close to the wall for vertical pipes, is relatively low.
In the upper part of an inclined pipe and in the core region of the vertical pipe, a
transitional region between the near wake and the far wake domains can be identified.
This region is characterized by a local minimum in the void fraction. The pattern of
the void fraction distribution was related to the flow field observed behind elongated
bubbles. The average void fraction values over the whole liquid slug were compared
to some results available in the literature and to the predictive model of Barnea and
Brauner (1985). Based on this model, a method was suggested to predict the high
voidage value in the near wake region of elongated bubbles. The overall mean void
fraction in the liquid slug ᾱs is given in Fig. 24 as a function of the mixture velocity
Um , for various pipe inclinations (Barnea et al., 2013). The overall void fraction was
obtained in this study by integration of the data accumulated in every cross-section,
within every liquid slug over the whole volume of the liquid slug.
At low mixture velocities (Um < 0.6 m/s) and at low pipe inclination angles (up
to β = 30◦ ), the liquid slug is almost free of dispersed bubbles (ᾱs ≈ 0). As the
mixture velocity increases, the average void fraction over the liquid slug increases
as well. Note that at higher pipe inclination angles, the average void fraction is
not zero, even at low values of Um . It was found that for identical values of Um ,
but different values of ULS and UGS , slightly higher values of ᾱs are obtained for
higher values of UGS . The mean void fraction ᾱs depends on both Um and the pipe
inclination angle β. Higher values of Um lead to stronger turbulent fluctuations in
the liquid slug that maintain higher void fraction in the slug body. The increase in the
inclination angle for a given value of Um lead to a higher shear rate between the wall
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 68

68 Modeling of Gas Liquid Flow in Pipes


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 24. Comparison of experimental data with predicted values of ᾱs (Barnea et al., 2013).
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 69

Hydrodynamic Models Based on Flow Patterns 69

jet at the entrance to the liquid slug from the film and thus to a stronger detachment
of dispersed bubbles from the elongated bubble’s tail.
The experimental results of Barnea et al. (2013) are juxtaposed in Fig. 24
with values available in literature. Symbols represent experimental data obtained
in different studies. The solid line represents the predicted values of ᾱs (Barnea
and Brauner, 1985) which are based on the theoretical bubble-slug transition lines
(Barnea, 1987). The dashed line employs the measured superficial velocities at the
slug-dispersed bubbles transition line as obtained by Barnea et al. (1985) to calculate
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

the corresponding values of ᾱs = UGS /(UGS + ULS ) and use those values to predict
the void fraction within the liquid slug.
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

In view of inevitable differences in the geometry of various experimental setups


and in experimental conditions, there is a reasonable agreement between the various
experimental results. Satisfactory agreement is obtained between the trend predicted
by the simplified model and the experimental results. Note also that the prediction
of ᾱs in the liquid slug is better when the experimental slug-bubble transition
boundary is used due to the intrinsic inaccuracy of the theoretical transition curve.

Prediction of ᾱ s
A model to predict the overall mean void fraction within the liquid slug ᾱs , was
suggested by Barnea and Brauner (1985). The model is based on the assumption
that the gas in the liquid slug behaves as dispersed bubbles on the verge of transition
to slug flow. The void fraction that the liquid slug can accommodate as dispersed
bubbles is to be determined from the balance between break up due to shear forces
in turbulent flow and coalescence due to gravity and/or surface tension. Whenever
coalescence dominates, agglomeration of small bubbles occurs, leading to formation
of elongated bubbles, which are separated by aerated liquid slugs. On the other
hand, with increasing turbulence level, breakage forces become dominating, leading
to formation of dispersed bubbles. This very same balance also determines the
transition boundary between dispersed bubbles and slug flow (Barnea, 1987). Since
the mixture velocity essentially determines the turbulence level, the liquid slug at
any flow rate is assumed to accommodate the same ᾱs , as dispersed bubble flow with
the same mixture velocity on the bubble-slug transition boundary. Thus, curves of
constant Um represent the locus where the gas void fraction within the liquid slug
ᾱs , is constant and equal to the gas void fraction of the dispersed bubble pattern
at the transition boundary. This transition boundary may be estimated either by a
predictive model, such as Barnea (1987) or from experimental data.
Dispersed bubble flow usually appears at very high liquid flow rates and is
observed over the whole range of pipe inclinations. Under certain conditions,
however, small discrete bubbles also appear at low liquid flow rates. These bubbles
are only observed in vertical and off-vertical (inclination angles exceeding about 55◦ )
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 70

70 Modeling of Gas Liquid Flow in Pipes

flows in pipes of relatively large diameter (D > 0.05 m). When these two conditions
are satisfied, the slug-bubbly flow transition line corresponds to a critical value
of = 0.25. For more details, see Barnea (1987) and references therein.
The predictive model of Barnea and Brauner (1985) refers to the void fraction
averaged over the whole body of the liquid slug ᾱs . A similar theoretical approach
is now applied to predict the void fraction specifically in the near wake of a
Taylor bubble. According to the model, the gas holdup on the dispersed bubble-
slug transition line is the maximum holdup that the liquid slug can accommodate
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

at a given turbulence level, which is determined by the mixture velocity Um . The


turbulence level in the body of the liquid slug is determined by the power dissipation
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

per unit mass, ε, in the pipe:


dP Um 4 Um 2
ε= =τ = f Um3 (130)
dx ρ D ρ D
Here dP/dx is the pressure gradient in the pipe; ρ, the mean fluid density and τ is
the wall shear stress. However, the power dissipation in the near wake region, εnw ,
is much higher and can be estimated as
εnw τnw u 2
= ∼ nw (131)
ε τ u 2
The characteristic ratio of the turbulent kinetic energy in the near wake of the Taylor
bubble to that in the far wake is O(102 ) (Shemer et al., 2005; 2007). The equivalent
value of Um,eq , that represents the power dissipation in the near wake region, can be
estimated as following:
εnw 2 u 2 2 3
εnw = ε ∼ f Um3 nw = f Um,eq (132)
ε D u 2 D
 3
1
u 2
Um,eq = nw
Um ∼ 5Um (133)
u 2
Substitution of Um,eq instead of Um in model of Barnea and Brauner (1985) provides
reasonable estimates of the average high values of αs in the near wake region as
presented in Barnea et al. (2013).

4.5.4. Slug length


Slug lengths and elongated bubble lengths are not deterministic values even for
“developed slug flow”. Most models that deal with steady slug flow assume
constant slug length (Dukler and Hubbard, 1975; Fernandes et al., 1983; Taitel
and Barnea, 1990). However the knowledge of the time average values is not always
sufficient and information on the distribution of the slug lengths and the elongated
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 71

Hydrodynamic Models Based on Flow Patterns 71

bubble lengths are often essential. Of particular importance is the maximum possible
slug length, since slug catchers design depends on the longest encountered slug
length and not necessarily on the average one. It is also important to know the
development of the slug length distribution along the pipe.
Slug length and slug frequency are inter-related properties and are often used
alternatively. Experimental observations for air–water systems in upward vertical
and horizontal flows suggest that the average stable liquid slug length is relatively
insensitive to the gas and liquid flow rates and depends mainly on the pipe diameter.
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

The average slug length has been observed to be about 15–40 pipe diameters for
horizontal flow (Dukler and Hubbard 1975; Nicholson et al., 1978; Fabre and Linè,
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

1992; Nydal et al., 1992; Andreussi et al., 1993; Felizola and Shoham, 1995) The
measured mean liquid slug length for vertical flow lies in the range of 8–25 pipe
diameters (Moissis and Griffith, 1962; Moissis, 1963; Akagawa and Sakaguchi,
1966; Fernandes, 1981; Barnea and Shemer, 1989; van Hout et al., 1992).
In the case of horizontal and slightly inclined flow, the slug frequency is
determined by the bridging of the air passage near the transition from stratified to slug
flow (Taitel and Dukler, 1977). For relatively low flow rates, the slugs are generated
at low frequency causing long liquid slugs near the entrance, which propagate
downstream. However, generally, short (high-frequency) slugs are formed at the
entrance of the pipe. These slugs are usually unstable. Shedding of liquid at the rear
of the liquid slug seems to be larger for short slugs. As a result, an elongated bubble
behind a short slug moves faster and overtakes the bubble ahead of it (Moissis and
Griffith, 1962). The bubble and the corresponding liquid slug merge in this process,
decreasing the slug frequency. The merging process continues until the liquid slug is
long enough to be stable, namely when the trailing bubble is unaffected by the wake
of the leading one. This occurs when the velocity profile at the rear of the liquid slug
can be considered fully developed (Moissis and Griffith, 1962; Taitel et al., 1980;
Barnea and Brauner, 1985; Dukler et al., 1985). Moissis and Griffith (1962), Taitel
et al. (1980), and Barnea and Brauner (1985) simulated the mixing process between
the film and the slug by a wall jet entering a large reservoir. It was suggested that
a developed liquid slug length is equal to the distance at which the jet has been
absorbed by the liquid. Using this approach, a value of 16D was obtained for the
minimum liquid slug length in vertical upflow and a value of 32D was obtained in
the case of horizontal flow. Dukler et al. (1985), developed a model in which the
minimum stable slug length is determined by the re-establishment length of the wall
boundary layer of the liquid slug. They found that the minimum stable slug length
is of the order of 20D. Although the final result of Dukler et al. (1985) is similar
to that of the previous analysis, the centerline velocity in their model increases
with the distance from the leading bubble contrary to experimental observation.
Shemer and Barnea (1987) detected the velocity field in the wake of an elongated
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 72

72 Modeling of Gas Liquid Flow in Pipes

bubble in vertical and horizontal flow by using the hydrogen bubble technique and
utilized the results for estimating the minimum stable slug length which is of the
order of 20D.
All the above approaches provide an estimation of the average stable slug length
and tell nothing about the slug length distribution and the maximum possible slug
length. As is well known, the slug length is widely dispersed around its average
(Fabre and Linè, 1992).
Brill et al. (1981), based on data from the Prudhoe bay field, were the first to
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

suggest that the slug length distribution follows a log-normal distribution for large-
diameter pipes. Dhulesia et al. (1991) investigated the dependence of the liquid slug
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

distribution in large pipe diameter (D = 0.2 and 0.4 m), on both the pipe diameter
and the position along the pipe. It was found that the slug length and its standard
deviation increase with pipe diameter. Nydal et al. (1992) measured the statistical
distributions of some slug characteristics in an air–water horizontal system. They
showed that the cumulative probability density function of the measured slug
lengths fits well a log-normal distribution. Bernicot and Drouffe (1989) proposed a
probabilistic approach for slug formation at the entrance of a horizontal pipe. They
also modeled the evolution of the length distribution by an individual equation for
each slug. Their approach is based on the concept that shedding for short slugs is
greater than that for long slugs. Saether et al. (1990) analyzed data from different
horizontal two- and three-phase pipeline systems and concluded that the liquid
slug length distribution obeys fractal statistics. Dhulesia et al. (1991) used a 1-D
Brownian motion with drift theory to obtain the slug length distribution. van Hout
et al. (1992) measured the slug length distribution in upward vertical flow and found
that the ratio between the standard deviation and the average is within 20–40%.
Barnea and Taitel (1993) developed a simplified model that computes the
evolution of the slug length distribution at any position along the pipe. The model
assumes that short randomly distributed slugs are formed at the entrance of the
pipe. The model tracks each individual bubble as it propagates along the pipe. The
propagation velocity of a trailing bubble is related to the maximum instantaneous
liquid velocity ahead of it. This velocity decreases with the distance from the leading
bubble. Bubbles behind short slugs travel faster than bubble behind long slugs. Thus,
the propagation velocity of each bubble depends on the separation distance from the
bubble ahead of it, such that Ut,i = f ( s,i). Trailing bubbles that are faster than the
leading ones will overtake the leading bubbles. The merging process results in
decreasing the frequency of the slugs along the pipe and increasing the lengths of
both elongated bubbles and liquid slugs. The process of overtaking is terminated,
once all of the slugs are long enough such that the velocity profile at the back of
the slugs is fully developed and all the bubbles propagate at the same velocity. The
translation velocity of a bubble Ut,i as a function of the length of the liquid slug ahead
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 73

Hydrodynamic Models Based on Flow Patterns 73

of it should be given as an input relation. The model results show that the evolution
of the length distribution along the pipe, the fully developed distribution, the average
slug length and the standard deviation are not sensitive to the slug length distribution
at the pipe entrance. The slug length distribution in the developed region seems to
follow approximately a log-normal shape. The model also provides information on
the length of the entry region needed to establish fully developed slug flow.
As has been mentioned, the input relation to the model is the instantaneous
translational velocity of the elongated bubble, as a function of the instantaneous
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

liquid slug length ahead of it. Moissis and Grifith (1962) measured the rise velocity
of a trailing bubble as a function of the separation distance from the leading one.
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

They suggested the following correlation:


 
Ut s
= 1 + a exp b (134)
U∞ D
where a and b are constants, U∞ is the translational velocity behind an infinity
long liquid slug and s is the separation distance between two consecutive bubbles.
Pinto and Campos (1996) investigated the coalescence of pairs of Taylor bubbles
rising in stagnant liquid covering a wide range of liquid viscosities. They correlated
the instantaneous bubble velocity as a function of the separation distance by linear
relations. Their research was extended by Pinto et al. (1998; 2001) for flowing
liquid in the laminar and in the turbulent regimes. Data on the rise velocity of Taylor
bubbles in air–kerosene as a function of the separation distance were presented in
Hasanein et al. (1996). Aladjem Talvy et al. (2000) carried out experiments with two
consecutive air bubbles injected into a vertical pipe of stagnant water and flowing
water and determined the translational velocity of the trailing bubble as a function
of the separation distance from the leading one.
Cook and Behnia (2000) compared experimental liquid slug length distribution
at 10 m from the inlet to distributions computed by the model of Barnea and Taitel
(1993) for a 50 mm pipe. The experiments and model compared well. Costigan
and Whalley (1997) measured liquid slug length distribution in a vertical 0.032 m
diameter pipe at about 6 m from the inlet. They concluded that the bubble and slug
lengths are approximately normally distributed. Hasanein et al. (1996) measured
the liquid slug length distribution for air–kerosene mixture at 2 m from the inlet
of a 0.0254 m vertical pipe. The general shape and statistical parameters of the
distributions compared well to the prediction model of Barnea and Taitel (1993).
van Hout et al. (2001; 2003) measured the evolution of continuous slug flow
for two pipe diameters, for the whole range of upwards pipe inclination and for a
wide range of liquid and gas flow rates, using optical fiber probes. The measuring
technique enable one to measure the instantaneous velocities of nose and tail of
each elongated bubbles, simultaneously with the slug length ahead of it. The liquid
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 74

74 Modeling of Gas Liquid Flow in Pipes

slug and elongated bubble lengths distributions, together with the dependence of
the elongated bubble velocity on the liquid slug ahead of it are presented at various
locations along the pipe.
Figure 25 is an example of the ensemble averaged instantaneous translational
velocities of the trailing bubble nose as a function of the separation distance from
the leading bubble, together with the slug length distribution for three inclination
angles and at four different locations along the pipe.
The accumulated data of the instantaneous bubble velocity normalized by the
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

Nicklin’s velocity Uinst /UNick , as a function of the liquid slug length ahead of the
bubble was averaged over different measurement positions for each flow rate and
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

inclination angle. The results for D = 0.024 m are presented in Fig. 26. The
data for different flow rates collapse on a single curve and approach the value
predicted by Nicklin’s correlation at sufficiently large separation distances. The
data were best fitted adopting the format suggested by Moissis and Grifith (1962)
(Eq. (134)). The correlations for the various pipe inclinations are shown in the
figure.
van Hout et al. (2001; 2003) used these correlations in the Barnea and Taitel
(1993) model for slug length distribution. They compared the model predictions
for the evolution of slug length distribution along the pipe with their experimental
results for vertical and inclined slug flow. In general, the shape, the mean values and
the development of the slug length distribution along the pipe are well compared.
Figure 27 demonstrates the variation of the sampled ensemble size N (van Hout
et al., 2001). The values of N are normalized by the corresponding ensemble size
at the pipe exit. This variation represents the coalescence rates, which appears to be
essentially independent of the pipe diameter and flow conditions. The coalescence
rate defines the extend of the entrance length of the slug which is about 100D.
As has been mentioned, the developing region of continuous slug flow is
governed by the interaction between consecutive elongated bubbles. The effect
of the leading bubble on the trailing one is manifested by the unsteady velocity
field, induced by the wake of the leading bubble. The trailing bubble shape and its
translational velocity are strongly affected by the unsteady velocity field in front
of the bubble, as the bubble tip follows the instantaneous local maximum velocity
ahead of it. This prompted the need for a deeper insight into the wake region of
single elongated bubbles moving in stagnant and flowing liquid. Comprehensive
studies of the flow field of the wake behind a Taylor bubble in vertical pipes were
carried out in both laminar and turbulent background flows using PIV (van Hout
et al., 2002a; Gulitski, 2005; Nogueira et al., 2006; Shemer et al., 2007). Ensemble
averaged velocities in the frame of reference moving with the Taylor bubble, as well
as the characteristic turbulent quantities were presented.
July 10, 2015 7:49
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in


by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Hydrodynamic Models Based on Flow Patterns

b1881-set-I-v1-ch04
Figure 25. Instantaneous translational velocity as a function of separation distance and the liquid slug length distribution based on the instantaneous
velocity at various positions along the pipe and for different inclination angles (van Hout et al., 2003). D = 0.024 m; ULS = 0.01 m/s; UGS = 0.41 m/s.

75

page 75
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 76

76 Modeling of Gas Liquid Flow in Pipes


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 26. Averaged instantaneous translational velocity as a function of separation distance for
different inclination angles and various flow rates. D = 0.024 m (van Hout et al., 2003).

4.5.5. Taylor bubble bottom oscillation


Polonsky et al. (1999a) observed an oscillatory motion of the Taylor bubble bottom
while propagating along the pipe, contrary to the bubble nose that retains a permanent
shape. The frequency and the amplitude of these oscillations were studied as a
function of bubble length and water flow rate.
Power spectra of the bubble tail oscillations in stagnant liquid are presented in
Fig. 28 for a number of bubble lengths. The resonant character of the spectra is
clearly seen, and the dominant frequency of the oscillations can be easily identified.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 77

Hydrodynamic Models Based on Flow Patterns 77


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 27. Variation of the ensemble size along the pipe (van Hout et al., 2001).

Figure 28. Spectra of bubble bottom oscillations (Polonsky et al., 1999a).

The amplitude of the bubble oscillations is apparently growing with the bubble
length, while the dependence of the frequency of those oscillations on the bubble
length is less pronounced.
The measured values of the dominant frequency of those oscillations are
summarized in Fig. 29. The dominant frequency seems to be essentially independent
of the liquid flow rate. It increases somewhat with the bubble length, and seems to
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 78

78 Modeling of Gas Liquid Flow in Pipes


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 29. Dominant frequency of the bubble bottom oscillations for various liquid velocities
(Polonsky et al., 1999a).

remain constant for longer bubbles. In most cases, the dominant frequency is about
4 Hz, and it does not exceed 5 Hz.
Polonsky et al. (1999a) proposed that the movement of the bubble bottom can be
modeled as that of a gas–liquid interface in a cylindrical container. Similar sloshing
movement of the liquid surface can easily be observed in a glass of water, after it
has been declined. In the frame of reference moving with the bubble bottom, such
standing wave on the free surface can be described in a linear approximation using
the potential flow theory. The general solution of the Laplace equation

∇2 = 0 (135)

for the velocity potential , in cylindrical coordinates (r, θ, x) with the origin at
the center of the bubble bottom is

(r, θ, x, t ) = As Js (kr)ekx + i(ωt + ε) (136)

where ω is the frequency of oscillations, k is the wave number and ε is the arbitrary
phase angle. In Eq. (136), Js is the Bessel function with the index s, and As is the
arbitrary amplitude of the oscillations (Abramowitz and Stegun, 1964). The standard
linearized kinematic and dynamic boundary conditions at the free surface can be
written as

∂ξ ∂

=− (137)
∂t ∂ x
x=0
 
∂ σ 1 1
− gx = − + (138)
∂t ρ R1 R2
where ξ is the instantaneous surface elevation, R1 and R2 are the principal radii
of curvature, and σ is the surface tension coefficient. This leads to the well-known
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 79

Hydrodynamic Models Based on Flow Patterns 79

dispersion relation for gravity–capillary waves


σ 3
ω2 = gk + k (139)
ρ
Assuming that the radial velocity on the bubble circumference is zero (rigid
boundary):

∂

=0 (140)
∂r
r=R
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

The dominant lowest mode is obtained from the equation J1 (k R) = 0. For a pipe
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

of R = 12.5 mm (experimental condition), Eqs. (139)–(140) yield a dominant first


antisymmetric mode with a frequency of 6.5 Hz. Neglecting surface tension results in
a slightly lower value of 6.05 Hz. The calculated dominant frequency is higher than
the experimentally measured values in Fig. 29. This discrepancy may result from
a number of reasons. The crudest assumption, however, adopted in the simplified
analysis is the non-penetrating boundary condition (Eq. (140)). Actually, the bubble
is surrounded by a circular wall jet with velocities notably exceeding those of the
moving bubble. Due to the complexity of this flow with an unknown free surface
and a shear flow in the circular jet, an exact solution can not be easily obtained. The
capability of the fast wall jet to consume the radial mass flux results in non-zero
radial velocity at r = R. Hence, the more appropriate equation for the determination
of the lowest eigenmode is

∂

= const. (141)
∂r
r=R
so that J1 (k R) = B, where B is in the range 0 < B < 0.5, but its exact
value cannot be specified. For example, the choice of B = 0.3 corresponds to a
dominant antisymmetric mode with a frequency of 4.75 Hz, close to the experimental
observations.
The adopted simplified linear potential model can thus grasp only the general res-
onant character of the observed oscillations and explain the apparent independence
of the resonant frequency on the bubble length and flow rate. It is, however, incapable
of providing any information regarding the amplitude of those oscillations, which
is determined by the complicated shear flow in the circumference of the bubble.
Both the bubble length and the liquid flow rate contribute to the increase in the shear
stress in the liquid film.
Liberzon et al. (2006) presented an experimental investigation of short capillary
waves observed on the air–water interface of Taylor bubbles rising in vertical pipes.
The excitation of those waves was attributed by the authors to bubble bottom
oscillations, which serve as an effective wave maker. The interfacial waves that are
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 80

80 Modeling of Gas Liquid Flow in Pipes

excited by the bottom oscillations and propagate upward along the bubble interface
are strongly affected by the downward water current in the liquid film that creates
a Doppler shift in the wave frequency (Peregrine, 1976; Lai et al., 1989). Since
the film velocity varies along the Taylor bubble, the length of the surface waves
varies accordingly and the waves become longer, when they approach the nose of
the bubble, where the downward film velocity is slower.
To analyze this phenomenon, consider two moving frames of reference. In the
first frame of reference which travels upward with the bubble velocity Ut , x denotes
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

the vertical coordinate relative to bubble bottom location at the pipe axis. In this
frame of reference, the liquid downward velocity within the film is Urel = Ufilm +Ut ;
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

where for simplicity the downward liquid film velocity Ufilm is considered to be
uniformly distributed in the radial direction. The local values of Ufilm (x) and the film
thickness δ(x) in the frame of reference moving with the bubble, can be calculated
using the Barnea (1990) model. The second frame of reference moving with the film
velocity, x , is defined, so that the mean liquid velocity in this frame of reference
vanishes. The relation between the vertical coordinates in these frames of reference
is therefore given by x = x − Urel (x)t.
A monochromatic wave excited by bubble bottom oscillations with the radian
frequency ω = 2π f , f being the wave-maker frequency, that propagates along the
bubble interface, can be described in a frame of reference moving with the bubble
as η = f (kx − ωt). The same wave in the frame of reference moving downwards
with the liquid film can be presented as η = f (kx − ω t). Note that contrary to the
frequency ω , the wave number k is invariant with respect to the change of frames of
references and is thus identical to that seen by an outside observer. Comparing the
wave presentations in both frames of reference and invoking the relation between
the vertical coordinates x and x yields

kx + kUrel (x)t − ωt = kx − ω t (142)

It can be easily shown that the effect of gravity on a nearly vertical interface is
negligible for waves under consideration that have lengths of the order of few mm.
The radian wave frequency ω is therefore related to the wave number k by the
dispersion relation appropriate for pure capillary waves over a liquid with surface
tension coefficient σ and density ρ

ω = ω + kUrel = (k 3 /σρ)tanh(kh) (143)

Moreover, in spite of the fact that the liquid film is quite narrow, the deep water
condition tanh(kh) = 1 is satisfied with reasonable accuracy for all wave lengths
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 81

Hydrodynamic Models Based on Flow Patterns 81


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Figure 30. Comparison of the measured lengths of interfacial waves with calculations in stagnant
water (Liberzon et al., 2006).

observed in the present study. Finally, the relation between the wave length λ =
2π/k, Urel , and the wave-maker frequency, f , is obtained from Eq. (143) as a
function of the distance from the bubble bottom x,

1 σ 1
= f + Urel (x) (144)
λ(x) ρλ(x) λ(x)
Figure 30 shows the variation of the ensemble-averaged interfacial wave length
along Taylor bubbles rising in stagnant water in three pipes (D = 14, 26, and
44 mm). Since the length of the waves is predominantly determined by the Doppler
shift induced by the current in the liquid film, the waves that are excited at the
bubble bottom, where the current velocity is at its maximum, are short. The wave
lengths then increase gradually as they propagate upward along the bubble, where
the film becomes wider and the current velocity decreases accordingly. Since the
film velocity is larger for larger pipe diameter, the wave length at a given distance
from the bubble nose decreases with increasing D. The measured data in stagnant
water agree well with the calculations.
The observed wave lengths are short near the bubble bottom, increasing toward
the bubble nose with decreasing film velocity. For longer bubbles, the generated
waves are too short and decay fast, explaining the appearance of waves on short
Taylor bubbles only. The downward film velocity increases with pipe diameter and
with water flow rate, resulting in excitation of shorter interfacial waves.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 82

82 Modeling of Gas Liquid Flow in Pipes

5. Severe Slugging
In the previous chapter, hydrodynamic “steady” slugs as well as undeveloped slugs
were considered. Slug flow initiation, growth and development into continuous
steady slug flow in pipelines are transient processes, which lead to slug length
distribution even in the developed zone of the pipe. Temporary slugs can also be
generated as a result of a sudden increase of gas or liquid flow rate when the flow
in the pipe is stratified (Taitel et al., 1978). In this case, a long slug is generated and
pushed out by the gas behind it.
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

In hilly terrain pipelines with varying inclinations, slugs appear at two length
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

scales: relative short hydrodynamic slugs or longer terrain induced slugs with length
determined from the pipeline topography. In this case, the liquid phase tends to
accumulate at the lower valleys to form long liquid bridges that block the gas
passage. The gas upstream is compressed, while long liquid slugs grow in the valleys.
Eventually the long slugs are blown out from one pipeline section to the next, due
to spontaneous expansion of the gas upstream. This results in a complex transient
flow with significant fluctuations in the outlet liquid and gas flow rates. This process
is also termed severe slugging.
Severe slugging is most commonly associated with a pipeline/riser system often
found in offshore oil production facility. Liquid accumulates at the bottom of the
riser until sufficient pressure is generated behind it, to push the liquid over the top
of the riser, overcoming the static head.
A general solution applied to hilly terrain pipelines is quite complex and it
is usually treated by numerical methods. De Henau and Raithby (1995a; 1995b)
developed one-dimensional transient two-fluid model for the simulation of slug
flow in a pipeline and validated it by comparing the model predictions to available
numerical and experimental data. De Henau and Raithby (1995c) performed an
experimental study of terrain induced slugging in a laboratory scale pipeline, made
of two uphill and two downhill sections and compare the results to their numerical
model prediction. Kjeldby et al. (2011; 2013), and Kjeldby and Nydal (2013)
proposed a Lagrangian slug tracking model for a solution of terrain flow. Yet, they
applied and verified the analysis based on data for a line-riser system. Malekzadeh
et al. (2012a; 2012b and 2012c) also performed severe slugging experiments in
a line-riser system and reported on five different types of flow behavior. They
used the dynamic simulation tool of OLGA for comparison with their experimental
observations.
Note that when treating terrain flow, one has to distinguish between low flow
rates in which case the dips accumulate liquid, resulting in terrain induced slugging
and high flow rates, where the effect of inclination changes is relatively minor and
hydrodynamic slugs flow along all sections. For example, in the case of annular flow,
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 83

Hydrodynamic Models Based on Flow Patterns 83

the inclination angles have little or no effect on the flow characteristics. For slug
flow, there is some effect on the distribution of slug lengths, as slugs tend to elongate
when passing through a dip and shrink when they pass through a top. These kinds
of hydrodynamic slugs in hilly terrain were treated by Barnea and Taitel (1993) and
Al-Safran et al. (2005).

5.1. The severe slugging cycle


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

The simple case of terrain induced slugging (severe slugging) that consists of a single
pipeline and a riser, is a common occurrence in an offshore oil and gas production
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

and has been analyzed quite successfully for the case of low flow rates. The result
of this analysis is detailed here.
Figures 31(a)–(d) show a typical behavior of the severe slugging cycle. The
first step is the slug formation (Fig. 31(a)). Liquid entering the pipe accumulates at

(a) Slug formation (b) Slug movement into the separator

(c) Blowout (d) Liquid fallback

Figure 31. The severe slugging cycle.


July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 84

84 Modeling of Gas Liquid Flow in Pipes

the bottom of the riser, blocks the gas passage and causes the gas in the pipeline
to compress. x is the distance of the liquid penetration into the line, z is the liquid
height in the riser, Pp is the pipeline pressure, PS is the separator pressure. When
the liquid height in the riser z, reaches the top of the riser, z = h, the second step of
slug movement into the separator starts (Fig. 31(b)). When the gas which is blocked
in the pipeline reaches the bottom of the riser, the liquid slug is accelerated to high
velocity, owing to rapid expansion of the gas in the pipeline. This step is termed
“blowout” (Fig. 31(c)). In the last step, Fig. 31(d), the remaining liquid in the riser
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

falls back to the bottom of the riser and the process of slug formation starts again.
The variation of x, z and Pp as a function of time during the severe slugging
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

cycle is calculated as follows:


Since the severe slugging phenomenon is typical of low flow rates of liquid and
gas, the pressure is dominated by gravity and the frictional contribution is neglected.
The liquid is considered incompressible and the gas is assumed to behave as an ideal
gas. Mass inlet flow rates of the liquid and the gas are assumed to be constant. With
reference to Fig. 31(a), x(t) and z(t) can be calculated using the following relations:
Mass conservation of liquid and gas yields:
t
m L = m L,i + AULS ρL dt (145)
0
t
m G = m G,i + AUGS0 ρG0 dt (146)
0

where the subscript 0 indicates reference condition, m Li and m Gi are the initial
masses of the liquid and gas in the pipeline. The determination of these initial
values is discussed later.
The mass of liquid and gas can be written in terms of the values of x and z as
follows:

m L = ρL A(x + z) + (1 − α)ρL A( − x) (147)


PP PS + ρL g(z − x sin β)
m G = ρG VG = VG = ( − x)α A (148)
RT RT
where A is the pipe cross-sectional area, VG is the volume of the gas in the pipe, T
is temperature, R is the ideal gas constant, β is the pipe inclination angle and α is
the void fraction in the line.
Note that m Li and m Gi are given by (147) and (148) for x = xi and z = z i .
Equations (145)–(148) are used to solve for x, z and PP as a function of time,
provided the initial values of x and z (xi and z i ) and α, are known.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 85

Hydrodynamic Models Based on Flow Patterns 85

Substituting m G and m Gi from (148) in (146) yields for the gas



PS
+ (z − x sin β) ( − x)α
ρL g
t
PS RT
= + (z i − xi sin β) ( − xi )α + UGS0 ρG0 dt (149)
ρL g ρL g
0

Substituting m L and m Li from (147) in (145) yields for the liquid:


Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

t
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

z = z i − α(x − xi ) + ULS dt (150)


0

Equations (149) and (150) can be solved now for x(t) and z(t) which correspond
to the slug formation step (Fig. 31(a)). Once the slug reaches the top of the riser
(z = h), the solution for x(t) is obtained directly from (149) with z = h.
The initial values of xi and z i depend on the amount of liquid that stays in the
riser at the end of the blowout process (Figs. 31(c) and 31(d)). The blowout process
is usually a highly chaotic phenomenon and the prediction of the amount of the
liquid fallback is difficult. Schmidt et al. (1980) used an experimental correlation
to estimate the amount of fallback. Taitel (1986) assumed that the blowout process
is in the form of a single fast moving Taylor bubble. In this case the void fraction
left in the riser α , can be calculated on the basis of a long Taylor bubble. The
result shows that the α , is around 90%. Assuming that at the beginning of the
fallback, the pressure in the pipeline is PS , the falling liquid blocks the air passage
and the fallback is very fast, then one can calculate xi , z i and Pp using the following
relations:
Hydrostatic pressure:

PP = PS + ρL g(z i − xi sin β) (151)

Liquid mass balance requires:

αxi + z i = (1 − α )h (152)

while the compression of the gas in the pipeline follows the relation

PP = PS (153)
( − xi )
The void fraction α in the pipeline can be calculated using a steady state stratified
flow in an inclined pipeline (Taitel and Dukler, 1976). Furthermore, since the flow
of the liquid and gas is usually low (this is a prerequisite for severe slugging), the
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 86

86 Modeling of Gas Liquid Flow in Pipes

void fraction can be calculated as for the case of an open channel flow. In this case,
a momentum balance of shear stress and gravity on the liquid phase yields:

τL SL = ρL g AL sin β (154)

where

ρL UL2
τL = f L (155)
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

2
The friction factor fL can be calculated from the Moody diagram with the appropriate
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

hydraulic diameter. For smooth pipes, for example, the friction factor can be
calculated by:
 −n
4AL UL ρL
f L = CL (156)
SL µL

where CL = 0.046, n = 0.2 for turbulent flow and CL = 16, n = 1 for laminar flow.
The cross-sectional area AL and the wetting periphery SL are given in terms of the
equilibrium liquid level h L . Equation (154) can now be solved for the equilibrium
level h L . Once h L is given, the void fraction α can be calculated by:

AL
α =1− (157)
A
The calculation presented here provides sufficient means to calculate some of the
major parameters of severe slugging such as: x(t), z(t), the fluctuation of the pressure
in the line, PP (t), the slug length that enters the separator, the cycle time of the
process, etc.
In the following sections, the criteria for the initiation of severe slugging (the
blowout step) is analyzed.

5.2. Boe’s criterion for severe slugging


The severe slugging pattern is typical of relatively low liquid and gas flow rates.
Thus, one condition for the existence of severe slugging is that the flow pattern in
the inclined pipeline is in the stratified flow pattern. For the determination of the
above condition, one needs to use a flow pattern map or any flow pattern prediction
methods (Taitel and Dukler, 1976; Barnea, 1987).
In addition to the above condition, the existence of a severe slugging cycle
requires that the liquid should penetrate into the pipeline, namely x > 0 (Boe, 1981).
This requirement is usually satisfied for relatively low gas flow rates. Referring to
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 87

Hydrodynamic Models Based on Flow Patterns 87

Fig. 31(a), the condition for x to stay at zero is when the increase of the pressure
owing to the addition of liquid into the riser is balanced by the increase in the pipeline
pressure due to the addition of gas.
The increase of pressure owing to the addition of liquid is:
dPP dz
= ρL g = ρL gULS (158)
dt dt
The increase of pressure owing to the addition of gas is:
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

dPP ṁ G UGS0 ρG0 PP UGS


= RT = RT = (159)
α α
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

dt VG
Equating the RHS of (158) and (159) yields the transition boundary proposed by Boe
between the severe slugging pattern and a steady flow in the riser (usually bubbly
or slug flow):
PP ρG0 RT
ULS = UGS or ULS = UGS0 (160)
ρL g α ρL g α
where the subscript 0 indicates reference conditions. Setting x and xi to zero in
Eqs. (149) and (150) leads to the same condition as Eq. (160).
Equation (160) is shown by boundary A on Fig. 32 for a specific example
reported by Taitel et al. (1990). Note, however, that α is calculated while neglecting
the gas shear (Eq. (154)), thus the transition for high liquid flow rates is less
accurate.
Boe claimed that outside the region bounded by A, the flow is of steady state
nature (bubbly flow or slug flow in the riser), while inside this region severe slugging

Figure 32. Occurrence of severe slugging, air–water (Taitel et al. 1990). T = 20◦ C, Ps = 0.1 MPa,
D = 2.54 cm, = 14 m, h = 3 m, β = 5◦ .
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 88

88 Modeling of Gas Liquid Flow in Pipes

prevails. This claim, however, is not quite accurate. In fact, the Boe’s criterion may be
“violated” and one may get steady state flow within the region designated by Boe as
“severe slugging” and vice versa, one can get severe slugging in the region designated
by the Boe’s criterion as steady state flow. The occurrence of such “anomalies” is
discussed next.

5.3. The stability criterion


The stability criterion addresses itself to the blowout step of the severe cycle process.
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

As discussed earlier, the blowout process (Fig. 31(c)) is assumed to take place in the
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

form of a spontaneous expansion of the gas in the pipeline. Indeed, this is usually the
case. The criterion for determining the condition under which a vigorous blowout
will occur versus a quasi-equilibrium penetration of the gas is termed here, the
stability criterion (Taitel, 1986).
Assume that the cycle of severe slugging reaches the point at which the liquid
slug tail has just entered the riser. Assume a small disturbance y which carries the
liquid somewhat higher (see Fig. 31(c)) and that the disturbance is fast enough so
that the slow flow rate of liquid and gas in the line is ignored, while y changes.
The net force (per unit area) acting on the liquid in the riser is:

α
F = (PS + ρL gh) − [PS + ρL g (h − y)] (161)
α + α y
where α is the gas holdup in the gas cap penetrating the liquid column.
The first term on the RHS is the pipeline pressure driving force. The pressure
varies with y as a result of the expansion of the gas in the pipeline. The second term
corresponds to the back pressure force applied by the separator pressure and the
liquid column of density ρL and height (h − y). Note that for y = 0, the system is
in equilibrium and F = 0. α can be estimated on the basis of a slug flow model.
Note that α and α have values typically above 0.8. Their exact values only slightly
affect the results. The gas is assumed to expand isothermally following the ideal
gas law.
The liquid column is blown out of the pipe, if F increases with y. Thus the
condition for stability is
∂(F)
< 0 at y = 0 (162)
∂y
This leads to the following criterion for stability
PS α /α − h
> (163)
P0 P0 /ρL g
where P0 is the atmospheric, or reference pressure.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 89

Hydrodynamic Models Based on Flow Patterns 89

Equation (163) is shown in Fig. 32 by boundary B which divides the region


bounded by the Boe’s criterion into two sub regions. The region below line B is
unstable and the blowout process is vigorous. The region above B is characterized
by a quasi-equilibrium penetration of the gas into the liquid. Taitel et al. (1990)
showed that this penetration can end up either with steady flow in the riser or it
can develop into a cyclic operation. The latter is termed “quasi-equilibrium severe
slugging” (to be discussed in Section 5.4).
Criterion (163) can also be used outside Boe’s region, where a steady state flow
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

is assumed to take place (bubbly or slug flow in the riser). Indeed, it can be shown that
an unstable sub-region exists outside Boe’s region. In this region a “severe slugging”
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

process takes place as follows: Gas in the pipeline spontaneously expands into the
riser and a blowout occurs, followed by liquid fallback. Thereafter, gas continues to
penetrate into the riser and bubbles through it, while the liquid (mixture) level in the
riser z, rises toward the top of the riser. At the time the liquid level reaches the top of
the riser, a steady state is expected to ensue. However, because of the inherent lack
of stability, blowout reoccurs. This gives rise to a cyclic severe slugging process,
except that the liquid entering the separator is aerated and shorter than the riser
length, unlike the “classic” severe slugging. The criterion for the existence of severe
slugging under such conditions is obtained using Eq. (163) in which ρL is replaced by
¯ where 
ρ L , ¯ is the average liquid holdup in the riser under steady state conditions
(the gas density can be ignored).
The liquid holdup in the riser, assuming stable slug flow in the riser is:

¯ = UGS
1− (164)
C(UGS + ULS ) + Ud

Note that Eq. (164) is also valid for bubbly flow in the riser in which case C and Ud
are replaced by B and U0 .
The gas superficial velocity in the riser, corrected for the average pressure in
the riser, is given by:

P0
UGS = UGS0 (165)
¯
PS + ρL gh /2

Equations (164) and (165) yield the liquid holdup , ¯ in the riser at steady state. The
stability of this steady state can be evaluated by Eq. (163) using ρ ¯ L . The line of
marginal stability is shown in Fig. 32 by curve C. As can be seen, there is a definite
region in which one can obtain unstable steady state flow outside Boe’s region. As
a result the flow is cyclic similar to the severe slugging cycle. This cyclic behavior
is termed “unstable oscillations”.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 90

90 Modeling of Gas Liquid Flow in Pipes

5.4. Quasi-equilibrium severe slugging


The region above line B in Fig. 32, although found to be stable according to Eq. (163),
may behave in a cyclic fashion termed quasi-equilibrium severe slugging. In this
case, it is possible to calculate the variation of the liquid holdup along the riser and
the line pressure as a function of time. Also one can predict whether the system ends
up in a steady state or a cyclic operation.
The analysis begins at the point when the riser is full of liquid, and gas front is
just entering the bottom of the riser. Within the region A and above B, the flow is
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

assumed to be stable so that no vigorous blowout occurs and the gas front penetrates
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

into the riser. As a result, the hydrostatic pressure at the riser bottom decreases
causing an expansion of the gas in the pipeline. In this case, the mass flow rate of
gas into the riser ṁ G increases. Assuming ideal gas behavior, the instantaneous mass
flow rate into the riser can be calculated by:
α A dP P
ṁ G = ṁ Gin − (166)
RT dt
where ṁ Gin is the gas inlet flow rate.
The pressure in the pipeline (and at the bottom of the riser) is the hydrostatic
pressure exerted by the weight of the liquid column in the riser (the gas weight is
neglected). Designating the local liquid holdup in the pipe as , one obtains:
h
PP = PS + ρL g dy (167)
0

The gas that penetrates the bottom of the riser develops into bubbly flow or Taylor
bubble flow.
Its average density is,
h )h
P
ρ̄G = (1 − ) dy (1 − )dy (168)
RT
0 0

where  is the local liquid holdup in the riser and P(y) is the local pressure.
h
P(y) = PS + ρL g dy (169)
y

The superficial gas velocity entering the riser is


ṁ G
UGS = (170)
ρ̄G A
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 91

Hydrodynamic Models Based on Flow Patterns 91

The liquid holdup at the bottom of the riser is given by:


UGS
b = 1 − (171)
Ut
where Ut , the translational velocity, is claculated using Eq. (96). US , the mixture
velocity, is calculated at the riser inlet and it is assumed constant along the riser
though it is a function of time.
The local liquid holdup in the riser is determined by tracking the propagation of
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

the liquid holdup at the bottom of the riser with a velocity Ut . Thus, the local liquid
holdup is calculated by:
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

t −τ
(y, t) = b (τ ) for y = Ut dt (172)
0

where τ is the specified time of the liquid holdup at the bottom of the riser and t is
any desired time (t > τ ).
This formulation allows one to calculate the variation of the pipeline pressure,
the gas mass flow rate ṁ G into the riser as a function of time and the local
instantaneous liquid holdup in the riser (y, t). It is easy to obtain a solution using
an explicit Lagrangian numerical scheme.
At time t = 0, the riser is full of liquid,  = 1 and ṁ G = ṁ Gin . The average
density of the gas at this time is the inlet density. The gas superficial velocity is
given by Eq. (170) and the translational velocity is calculated by Eq. (96). The riser
is subdivided into small segments of length h and the time step t is calculated
using t = h/Ut .
After a time interval t, 1 (at the bottom of the riser, = b ) is given by
Eq. (171); the new pressure is given by Eq. (167); the new average gas density in
the riser is given by Eq. (168); and the new gas mass flow rate into the riser is given
by Eq. (166). Note that dPP /dt in Eq. (166) is approximated numerically by the
difference between the “new” and “old” pressure divided by t. Once the new ṁ G
is known, the new gas superficial velocity UGS is calculated from Eq. (170) along
with the new translational velocity Ut and the new time step t (t = h/Ut ). At
the next time step, the values of  j +1 are set equal to  j . 1 is calculated as before.
If the analysis yields a positive gas flow rate into the riser ṁ G , then steady state
is reached (above boundary D). If ṁ G becomes negative, penetration of liquid into
the pipeline takes place and the outcome is a cyclic process (below D).
Let x(t) be the distance of the liquid interface penetrating into the pipeline.
Under hydrostatic equilibrium the pipeline pressure at any time is:

¯ − x sin β) + PS
PP = ρL g(h (173)
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 92

92 Modeling of Gas Liquid Flow in Pipes

where ¯ is the average liquid holdup in the riser. A mass balance on the gas in the
pipeline requires that (see Eq. (149)):

PS 
¯
+ h − x sin β ( − x) α
ρL g
t
PS RT
= ¯
+ i h α + ṁ Gin dt (174)
ρL g AρL g
0
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

where i relates to the time, when ṁ G = 0 and penetration of liquid into the pipe
starts.
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Equation (174) can be solved for x as a function of time. For this purpose, the
average liquid holdup  ¯ should be known as a function of time. The variation of ¯
with time can be calculated as before, on the basis of the translational velocity Ut
from Eq. (96). The mixture velocity US is then calculated on the basis of the liquid
mass balance to yield:
dx
US = ULS − α (175)
dt
At time ti , x = 0 and US = ULS (ṁ G and UGS = 0). For time step t, the new values of
¯ x, US (approximating dx/dt numerically), Ut and the new time
 (z) in the riser, ,
step t, are calculated. As in the case of the severe slugging cycle, x increases to a
maximum and then recedes back to zero. When x = 0, the cyclic process is repeated.

5.5. Summary
The severe slugging that consists of one riser and one pipeline is perhaps one of the
simplest examples of slug flow under non-steady conditions. As is evident, even this
simple case is not at all trivial and presents quite a number of possible operating
conditions. A more detailed discussion and experimental verification of the present
theory is given by Taitel (1986) and Taitel et al. (1990). A summary of the results
are presented by a typical flow map as shown in Fig. 32.
This map contains four boundaries: A, the Boe’s criterion; B, the stability
criterion inside the Boe’s region; C, the stability criterion outside the Boe’s region
and D, the transition to steady flow inside the Boe’s criterion.
The Boe’s criterion (Eq. (160)) differentiates quite well between steady and
cyclic operations with two exceptions. At high liquid flow rates, a steady flow can
also exist within the “severe slugging” region predicted by the Boe’s criterion (above
boundary D). Also there is a region outside the Boe’s criterion which is in an
unsteady state and leads to unsteady oscillations (between boundaries C and A).
The stability criterion (Eq. (163)) is applied to the case of severe slugging (inside
the Boe’s region), where the riser contains only liquid (transition B) and to the case
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 93

Hydrodynamic Models Based on Flow Patterns 93

of steady flow of liquid and gas in the riser outside the Boe’s region (transition C).
Boundary B divides between the “classical” blowout and a quasi-equilibrium severe
slugging cyclic operations. Boundary C indicates when steady flow outside the
Boe’s criterion is not possible and one obtains unsteady oscillations. Transition D
demarcates between cyclic and steady state operation within the Boe’s region.

References
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

Abramowitz, M. and Stegun, I.A. (1964). Handbook of Mathematical Functions. U.S.


by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

National Bureau of Standards, Dover, New York.


Abdul-Majeed, G.H. (2000). Liquid slug holdup in horizontal and slightly inclined two-
phase slug flow. J. Petr. Sci. Eng. 27, pp. 27–32.
Agrawal, S.S., Gregory, G.A. and Govier, G.W. (1973). An analysis of horizontal stratified
two phase flow in pipes. Canadian J. Chemical Engineering 51, pp. 280–286.
Aladjem Talvy, C., Shemer, L. and Barnea D. (2000). On the interaction between
two consecutive elongated bubbles in a vertical pipe. Int. J. Multiphase Flow 26,
pp. 1905–1923.
Al-Safran, E., Sarica, C., Zang, H.Q. and Brill, J. (2005). Investigation of slug flow
characteristics in the valley of hilly — terrain pipeline. Int. J. Multiphase Flow 31,
pp. 337–357.
Al-Safran, E. (2009). Prediction of slug liquid holdup in horizontal pipes. J. Energy Resour.
Technol. 131, pp. 023001-1–023001-8.
Alves, I.N., Shoham, O. and Taitel, Y. (1993). Drift velocity of elongated bubbles in inclined
pipes-experimental and modeling. Chem. Eng. Sci. 48, pp. 3063–3070.
Akagawa, K. and Sakaguchi, T. (1966). Fluctuation of void ratio in two-phase flow (2nd
report, analysis of flow configuration considering the existence of small bubbles in
liquid slugs). Bull. JSME 9, pp. 104–110.
Akagawa, K., Hamaguchi, H., Sakaguchi, T. and Ikari, T. (1971). Studies on the fluctuation
of pressure drop in two-phase slug flow. Bull. JSME 14, pp. 447–454.
Andritsos, N. and Hanratty, T.J. (1987a). Influence of interfacial waves on holdup and
frictional pressure drop in stratified gas–liquid flows. AIChE J. 33, pp. 444–454.
Andritsos, N. and Hanratty, T.J. (1987b). Interfacial instabilities for horizontal gas–liquid
flows in pipelines. Int. J. Multiphase FIow 13, pp. 583–603.
Andreussi, P. and Bendiksen, K.H. (1989). An investigation of void fraction in liquid
slugs for horizontal and inclined gas–liquid pipe flow. Int. J. Multiphase Flow 15,
pp. 937–946.
Andreussi, P., Bendiksen, K.H. and Nydal, O.J. (1993). Void distribution in slug flow. Int.
J. Multiphase Flow 19, pp. 817–828.
Barnea, D. and Brauner, N. (1985). Holdup of the liquid slug in two-phase intermittent flow.
Int. J. Multiphase Flow 11, pp. 43–49.
Barnea, D., Shoham, O., Taitel, Y. and Dukler, A.E. (1985). Gas–liquid flow in inclined
tubes: fIow pattern transitions for upward flow. Chem. Eng. Sci. 40, pp. 131–136.
Barnea, D. (1987). A unified model for predicting flow pattern transitions in the whole range
of pipe inclination. Int. J. Multiphase Flow 13, pp. 1–12.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 94

94 Modeling of Gas Liquid Flow in Pipes

Barnea, D. and Shemer, L. (1989). Void fraction measurement in vertical slug flow:
Applications to slug characteristics and transitions. Int. J. Multiphase Flow 15,
pp. 495–504.
Barnea, D. (1990). Effect of bubble shape on pressure drop calculations in vertical slug flow.
Int. J. Multiphase Flow 16, pp. 79–89.
Barnea, D. and Taitel, Y. (1993). A model for slug length distribution in gas–liquid slug
flow. Int. J. Multiphase Flow 19, pp. 829–838.
Barnea, D., Roitberg, E. and Shemer, L. (2013). Spatial distribution of void fraction in
the liquid slug in the whole range of pipe inclinations. Int. J. Multiphase Flow 52,
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

pp. 92–101.
Barr, G. (1926). The air–bubble viscometer. Philos. Mag. 1, pp. 395–405.
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Bendiksen, K.H. (1984). An experimental investigation of the motion of long bubbles in


inclined tubes. Int. J. Multiphase Flow 10, pp. 467–483.
Bendiksen, K.H. (1985). On the motion of long bubbles in vertical tubes. Int. J. Multiphase
Flow 11, pp. 797–812.
Benjamin, T.B. (1968). Gravity currents and related phenomena. J. Fluid Mechanics 31,
pp. 209–248.
Bergelin, O.P. and Gazley, C. (1949). Co-current gas–liquid flow in horizontal tubes. Proc.
Heat Transf. and Fluid Mech. Inst. 29, pp. 5–18.
Bernicot, M. and Drouffe, J.M. (1989). Slug length distribution in two-phase transportation
systems. In: Cranfield, B. (ed.), Proc. Fourth International Conference on Multiphase
Flow, Nice, France, BHRA, pp. 485–493.
Boe, A. (1981). Severe slugging characteristics. Sel. Top. Two-Phase Flow. NTH, Trond-
heim, Norway.
Bonnecaze, R.H., Eriskine, W., Jr. and Oreskovich, E.J. (1971). Holdup and pressure drop
for two-phase slug flow in inclined pipelines. AIChE J. 17, pp. 1109–1113.
Brill, J. P., Schmidt, Z., Coberly, W.A., Herring, J.D. and Moore, D.W. (1981). Analysis of
two-phase tests in large-diameter flow lines in Prudhoe bay field. Society of Petroleum
Engineers Journal 271, pp. 363–378.
Brauner, N. and Ullmann, A. (2004). Modelling of gas entrainment from Taylor bubbles.
Int. J. Multiphase Flow 30, pp. 239–272.
Clift, R., Grace, J.R. and Weber, M.E. (1978). Bubbles, Drops and Particles. Academic
Press, New York.
Cohen, S.L. and Hanratty, T.J. (1968). Effects of waves at a gas–liquid interface on a turbulent
air flow. J. Fluid Mechanics 31, pp. 467–469.
Collins, R., De Moraes, F.F., Davidsom, J.F. and Harrison, D. (1978). The motion of a
large gas bubble rising through liquid flowing in a tube. J. Fluid Mech. 89, Part 3,
pp. 497–514.
Cook, M. and Behnia, M. (2000). Slug length prediction in near horizontal gas–liquid
intermittent flow. Chem. Eng. Sci. 55, pp. 2009–2018.
Costigan, G. and Whalley, P.B. (1997). Slug flow regime identification from dynamic
void fraction measurements in vertical air–water flows. Int. J. Multiphase Flow 23,
pp. 263–282.
Davies, R.M. and Sir Taylor, G. (F. R. S). (1949). The mechanics of large bubbles rising
through extended liquids and through liquids in tubes. Proc. R. Soc. London Ser. A 200,
pp. 375–390.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 95

Hydrodynamic Models Based on Flow Patterns 95

De Henau, V. and Raithby, G.D. (1995a). A transient two-fluid model for the simulation of
slug flow in pipelines — I, theory. Int. J. Multiphase Flow 21, pp. 335–349.
De Henau, V. and Raithby, G.D. (1995b). A transient two-fluid model for the
simulation of slug flow in pipelines — II, validation. Int. J. Multiphase Flow 21,
pp. 351–363.
De Henau, V. and Raithby, G.D. (1995c). A study of terrain-induced slugging in two- phase
flow pipelines. Int. J. Multiphase Flow 21, pp. 365–379.
Delfos, R. (1996). Experiments on air entrainment from a stationary slug bubble in a vertical
pipe. Ph.D. Thesis. Delft University of Technology, Delft, Netherlands.
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

Dhulesia, H., Bernicot M. and Deheuvels, P. (1991). Statistical analysis and modeling of
slug lengths. Proc. Fifth International Conference on Multiphase Production, Cannes,
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

France, BHRA.
Dukler, A.E. and Hubbard, M.G. (1975). A model for gas–liquid slug flow in horizontal and
near horizontal tubes. Ind. Eng. Chem. Fundam 14, pp. 337–347.
Dukler, A.E., Moalem-Maron, D. and Brauner, N. (1985). A physical model for predicting
the minimum stable slug length. Chem. Eng. Sci. 40, pp. 1379–1385.
Dumitrescu, D.T. (1943). Stromung an einer luftblase im senkrechten rohr. Z. Angew Math.
Mech. 23, pp. 139–149.
Fabre, J. and Linè, A. (1992). Modeling of two-phase slug flow. Annu. Rev. Fluid Mech. 24,
pp. 21–46.
Felizola, H. and Shoham, O. (1995). A unified model for slug flow in directional wells.
ASME J. Energy Resources Technology 117, pp. 7–12.
Fernandes, R.C. (1981). Experimental and theoretical studies of isothermal upward
gas–liquid flows in vertical tubes. Ph.D. Thesis. University of Houston, Houston,
Texas.
Fernandes, R.C., Semiat, R. and Dukler, A.E. (1983). Hydrodynamic model for gas–liquid
slug flow in vertical tubes. AIChE J. 29, pp. 981–989.
Fossa, M., Guglielmini, G. and Marchitto, A. (2003). Intermittent flow parameters from
void fraction analysis. Flow Meas. Instrum. 14, pp. 161–168.
Fréchou, D. (1986). Etude de l’écoulement ascendant à trois fluides en conduite verticale.
Thèse Inst. Natl. Polytech., Toulouse.
Goldsmith, H.L. and Mason, S.G. (1962). The movement of single large bubbles in closed
vertical tubes. J. Fluid Mechanics 14, pp. 42–58.
Gomez, L.E., Shoham, O. and Taitel, Y. (2000). Prediction of slug liquid holdup horizontal
to upward vertical flow. Int. J. Multiphase Flow 26, pp. 517–521.
Govier, G.W. and Omer, M.M. (1962). The horizontal pipeline flow of air–water mixtures.
Canadian J. Chemical Engineering 40, pp. 93–104.
Govier, G.W. and Aziz, K. (1972). The Flow of Complex Mixtures in Pipes. van Nostrand
Reinhold, New York.
Grace, J.R. and Clift, R. (1979). Dependence of slug rise velocity on tube Reynolds number
in vertical gas–liquid flow. Chem. Eng. Sci. 34, pp. 1348–1350.
Gregory, G.A., Nicholson, M.K. and Aziz, K. (1978). Correlation of the liquid volume
fraction in the slug for horizontal gas–liquid slug flow. Int. J. Muftiphase Flow 4,
pp. 33–39.
Guet, S., Decarre, S., Henriot, V. and Line, A. (2006). Void fraction in vertical gas–liquid
slug flow: Influence of liquid slug content. Chem. Eng. Sci. 61, pp. 7336–7350.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 96

96 Modeling of Gas Liquid Flow in Pipes

Gulitski, A. (2005). Experimental investigation of the flow field in the wake of a Taylor
bubble and its effect on the motion of a consecutive bubble in vertical slug flow. Ph.D.
Thesis. Tel Aviv University, Tel Aviv, Israel.
Hall, N.A. (1957). Thermodynamics of Fluid Flow. Longmans, Green, New York.
Harmathy, T.Z. (1960). Velocity of large drops and bubbles in media of infinite or restricted
extent. AIChE J. 6, pp. 281–288.
Hasan, A.R. and Kabir, C.S. (1986). Predicting multiphase flow behavior in a deviated well.
61st Annu. Tech. Conf., New Orleans, LA, SPE 15449.
Hasanein, H.A., Tudose, G.T., Wong, S., Malik, M., Esaki, S. and Kawaji, M. (1996). Slug
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

flow experiments and computer simulation of slug length distribution in vertical pipes.
AIChE Symposium Series Heat Transfer, Houston.
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Haywood, N.I. and Richardson, J.F. (1979). Slug flow of air–water mixtures in a horizontal
pipe: Determination of liquid holdup by gamma-ray absorption. Chem. Eng. Sci. 34,
pp. 17–30.
Henstock, W. and Hanratty, T.J. (1976). The interfacial drag and the height of the wall layer
in annular flow. AIChE. J. 22, pp. 990–1000.
Henderson, F.M. (1966). Open Channel Flow. Macmillan, New York.
Hewitt, G.F. and Hall-Taylor, N.S. (1970). Annular Two-Phase Flow. Pergamon Press,
Oxford.
Hoogendoorn, C.J. (1959). Gas–liquid flow in horizontal pipe. Chem. Eng. Sci. 9,
pp. 205–217.
Johannessen, T. (1972). A theoretical solution of the Lockhart and Martinelli flow model
for calculating two-phase flow pressure drop and holdup. Int. J. Heat Mass Transfer
15, pp. 1443–1449.
Kaji, R., Azzopardi, B.J. and Lucas, D. (2009). Investigation of flow development of
cocurrent gas–liquid vertical slug flow. Int. J. Multiphase Flow 35, pp. 335–348.
Kjeldby, T.K., Henkes, R.A.W.M. and Nydal O.J. (2011). Slug tracking simulation of
severe slugging experiments. World Academy of Science, Engineering and Technol-
ogy 78.
Kjeldby, T.K., Henkes, R.A.W.M. and Nydal O.J. (2013). Lagrangian slug flow modeling
and sensitivity on hydrodynamic slug initiation methods in a severe slugging case. Int.
J. Multiphase Flow 53, pp. 29–39.
Kjeldby, T.K. and Nydal, O.J. (2013). A Lagrangian three phase slug tracking framework.
Int. J. Multiphase Flow 56, pp. 184–194.
Kockx, J.P., Nieuwstadt, F.T.M., Oliemans, R.V.A. and Delfos, R. (2005). Gas entrainment
by a liquid film falling around a stationary Taylor bubble in a vertical tube. Int. J.
Multiphase Flow 31, pp. 1–24.
Lai, R.J., Long, S.R. and Huang, N.E. (1989). Laboratory studies of wave current interaction:
kinematics of the strong interaction. J. Geophys. Res. 94, pp. 16201.
Liberzon D., Shemer, L. and Barnea, D. (2006). Upward-propagating capillary waves on
the surface of short Taylor bubbles. Physics of Fluids 18, pp. 048103, 1–4.
Lockhart, R.W. and Martinelli, R.C. (1949). Proposed correlation of data for isothermal
two-phase, two-component flow in pipes. Chem. Eng. Prog. 45, pp. 38–49.
Malekzadeh, R., Henkes, R.A.W.M. and Mudde, R.F. (2012a). Severe slugging in a
long pipeline-riser system: Experiments and predictions. Int. J. Multiphase Flow 46,
pp. 9–21.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 97

Hydrodynamic Models Based on Flow Patterns 97

Malekzadeh, R., Belfroid, S.P.C. and Mudde, R.F. (2012b). Transient drift flux modeling of
severe slugging in pipeline-riser systems. Int. J. Multiphase Flow 46, pp. 32–37.
Malekzadeh, R., Mudde, R.F., Henkes, R.A.W.M. (2012c). Dual-frequency severe slugging
in horizontal pipeline-riser systems. J. Fluids Eng. 134, pp. 121301-1–127309-1.
Mao, Z. and Dukler, A.E. (1991). The motion of Taylor bubbles in vertical tubes. II.
Experimental data and simulations for laminar and turbulent flow. Chem. Eng. Sci.
46, pp. 2055–2064.
Moissis, R. and Griffith, P. (1962). Entrance effects in a two-phase slug flow. J. Heat Transfer
84, pp. 29–39.
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

Moissis, R. (1963). The transition from slug to homogeneous two-phase flows. J. Heat
Transfer 85, pp. 366–370.
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Mori, K. and Miwa, M. (2002). Structure and void fraction in a liquid slug for gas–liquid
two-phase slug flow. Heat Transfer — Asian Res. 31, pp. 57–271.
Nädler, M. and Mewes, D. (1995). Effects of the liquid viscosity on the phase distributions
in horizontal gas–liquid slug flow. Int. J. Multiphase Flow 21, pp. 253–266.
Nicklin, D.J., Wilkes, J.O. and Davidson, J.F. (1962). Two-phase flow in vertical tubes.
Trans.1st. Chem. Eng. 40, pp. 61–68.
Nicholson, M.K., Aziz, K. and Gregory, G.A. (1978). Intermittent two phase flow
in horizontal pipes: predictive models. Canadian J. Chemical Engineering 56,
pp. 653–663.
Nogueira, S., Reithmuller, M.L., Campos, J.B.L.M. and Pinto, A.M.F.R. (2006). Flow
patterns in the wake of a Taylor bubble rising through vertical columns of stag-
nant and flowing Newtonian liquids: an experimental study. Chem. Eng. Sci. 61,
pp. 7199–7212.
Nydal, O.J., Pintus, S. and Andreussi, P. (1992). Statistical characterization of slug flow in
horizontal pipes. Int. J. Multiphase Flow 18, pp. 439–453.
Orell, A. and Rembrand, R. (1986). A model for gas–liquid slug flow in a vertical tube. Ind.
Eng. Chem. Fundamen. 25, pp. 196–206.
Peregrine, D.H. (1976). Interaction of water waves and currents. Adv. Appl. Mech. 16, pp. 9.
Perez, V.H. (2007). Gas–liquid two-phase flow in inclined pipes. Ph.D. Thesis. University
of Nottingham, U.K.
Pinto, A.M.F.R. and Campos, J.B.L.M. (1996). Coalescence of two gas slugs rising in a
vertical column of liquid. Chem. Eng. Sci. 51, pp. 45–54.
Pinto, A.M.F.R., Coelho Pinheiro, M.N. and Campos, J.B.L.M. (1998). Coalescence of two
gas slugs rising in a co-current flowing liquid in vertical tubes. Chem. Eng. Sci. 53,
pp. 2973–2983.
Pinto, A.M.F.R., Coelho Pinheiro, M.N. and Campos, J.B.L.M. (2001). On the interaction of
Taylor bubbles rising in two phase co-current slug flow in vertical columns: Turbulent
wakes. Exp. Fluids 31, pp. 643.
Polonsky, S., Barnea, D. and Shemer, L. (1999a). Averaged and time-dependent character-
istics of the motion of an elongated bubble in a vertical pipe. Int. J. Multiphase Flow
25, pp. 795–812.
Polonsky, S., Shemer, L. and Barnea, D. (1999b). The relation between the Taylor bubble
motion and the velocity field ahead of it. Int. J. Multiphase Flow 25, pp. 957–975.
Reinecke, N., Petrisch, G., Boddem, M. and Mewes, D. (1998). Tomographic imaging of
the phase distribution in two-phase slug flow. Int. J. Multiphase Flow 24, pp. 617–634.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 98

98 Modeling of Gas Liquid Flow in Pipes

Saether, G., Bendiksen, K., Müller, J. and Froland, E. (1990). The fractal statistics of liquid
slug lengths. Int. J. Multiphase Flow 16, pp. 1117–1129.
Schmidt, Z., Brill, J.P. and Beggs, H.D. (1980). Experimental study of severe slugging in a
two-phase flow pipeline-riser pipe system. Soc. Pet. Eng. J. 20, pp. 407–414.
Shemer, L. and Barnea, D. (1987). Visualization of the instantaneous velocity profiles in
gas–liquid slug flow. Physicochem. Hydrodyn. 8, pp. 243–253.
Shemer, L., Gulitski, A. and Barnea, D. (2005). Experiments on the turbulent structure
and the void fraction distribution in the Taylor bubble wake. Multiphase Sci. Tech. 16,
pp. 1–20.
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

Shemer, L., Gulitski, A. and Barnea, D. (2007). On the turbulent structure in the wake of
Taylor bubbles rising in vertical pipes. Phys. of Fluids 19, pp. 035108, 1–13.
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

Shoham, O. and Taitel, Y. (1984). Stratified turbulent–turbulent gas liquid flow in horizontal
and inclined pipes. AIChE J. 30, pp. 377–385.
Singh, G. and Griffith, P. (1970). Determination of the pressure drop and optimum pipe size
for a two-phase slug flow in an inclined pipe. J. Eng. Ind. 92, pp. 717–726.
Stanislav, J.F., Kokal, S. and Nicholson, M.K. (1986). Intermittent gas–liquid flow in upward
inclined pipes. Int. J. Multiphase Flow 12, pp. 325–335.
Sylvester, N.D. (1987). A mechanistic model for two phase vertical slug flow in pipes.
J. Energy Resour. Technol. 109, pp. 206–213.
Taitel, Y. and Dukler, A.E. (1976). A model for prediction flow regime transitions in
horizontal and near horizontal gas–liquid flow. AIChE J. 22, pp. 47–55.
Taitel, Y. and Dukler, A.E. (1977). A model for slug frequency during gas–liquid flow in
horizontal and near horizontal pipes. Int. J. Multiphase Flow 3, pp. 585–596.
Taitel, Y., Lee, N. and Dukler, A.E. (1978). Transient gas–liquid flow in horizontal pipes,
modeling flow pattern transitions. AIChE. J. 24, pp. 920–935.
Taitel, Y., Barnea, D. and Dukler, A.E. (1980). Modelling flow pattern transitions for steady
upward gas–liquid flow in vertical tubes. AIChE J. 26, pp. 345–354.
Taitel, Y. (1986). Stability of severe slugging. Int. J. Multiphase Flow 12, pp. 203–217.
Taitel, Y., Vierkandt, S., Shoham, O. and Brill, J.P. (1990). Severe slugging in a pipeline-riser
system, experiments and modeling. Int. J. Multiphase Flow 16, pp. 57–68.
Taitel, Y. and Barnea, D. (1990). A consistent approach for calculating pressure drop in
inclined slug flow. Chem. Eng. Sci. 45, pp. 1199–1206.
van Hout, R., Barnea, D. and Shemer, L. (2001). Evolution of statistical parameters of
gas–liquid slug flow along vertical pipes. Int. J. Multiphase Flow 27, pp. 1579–1602.
van Hout, R., Shemer, L. and Barnea, D. (1992). Spatial distribution of void fraction within
the liquid slug and some other related slug parameters. Int. J. Multiphase Flow 18,
pp. 831–845.
van Hout, R., Gulitski, A., Barnea, D. and Shemer, L. (2002a). Experimental investigation of
the velocity field induced by a Taylor bubble rising in stagnant water. Int. J. Multiphase
Flow 28, pp. 579–596.
van Hout, R., Barnea, D. and Shemer, L. (2002b). Translational velocities of elongated
bubbles in continuous slug flow. Int. J. Multiphase Flow 28, pp. 1333–1350.
van Hout, R., Shemer, L. and Barnea, D. (2003). Evolution of hydrodynamic and
statistical parameters of gas–liquid slug flow along inclined pipes. Chem. Eng. Sci. 58,
pp. 115–133.
July 10, 2015 7:49 Encyclopedia of Two-Phase Heat Transfer and Flow I - 9.75in x 6.5in b1881-set-I-v1-ch04 page 99

Hydrodynamic Models Based on Flow Patterns 99

Viana, F., Pardo, R., Ynez., R., Trallero, J.L. and Joseph, D.D. (2003). Universal correlation
for the rise velocity of long gas bubbles in round pipes. J. Fluid Mech. 494, pp. 379–398.
Wallis, G.B. (1969). One Dimensional Two-Phase Flow. McGraw-Hill, New York.
Wallis, G.B., Richter, H.J. and Bharathan, D. (1978). Air–water countercurrent annular flow
in vertical tubes. Rep. EPRI NP-786. Electr. Power Res. Inst. Palo Alto, California, US.
Wallis, G.B., Richter, H.J. and Bharathan, D. (1979). Air–water countercurrent annular flow
in vertical tubes. Rep. EPRI NP-1165. Electr. Power Res. Inst. Palo Alto, California, US.
White, E.T. and Beardmore, R.H. (1962). The velocity of rise of single cylindrical air
bubbles through liquids contained in vertical tubes. Chem. Eng. Sci. 17, pp. 351–361.
Encyclopedia of Two-Phase Heat Transfer and Flow I Downloaded from www.worldscientific.com

Yan, K. and Che, D. (2011). Hydrodynamic and mass transfer characteristics of slug flow
in a vertical pipe with and without dispersed small bubbles. Int. J. Multiphase Flow 37,
by CHINESE UNIVERSITY OF HONG KONG on 10/15/15. For personal use only.

pp. 299–325.
Zhang, H.Q., Wang, Q., Sarica, C. and Brill, J.P. (2003). A unified mechanistic model
for slug liquid holdup and transition between slug and dispersed bubble flows. Int. J.
Multiphase Flow 29, pp. 97–107.
Zheng, D. and Che, D. (2006). Experimental study on hydrodynamic characteristics of
upward gas–liquid slug flow. Int. J. Multiphase Flow 32, pp. 1191–1218.
Zuber, N. and Findlay, J.A. (1965). Average volumetric concentration in two-phase flow
systems. J. Heat Transfer 87, pp. 453–468.
Zukoski, E.E. (1966). Influence of viscosity, surface tension and inclination angle on motion
of long bubbles in closed tubes. J. Fluid Mechanics 25, pp. 821–837.

You might also like