You are on page 1of 37

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/300569742

Two-Phase Pressure Drop

Chapter · October 2015


DOI: 10.1142/9789814623216_0023

CITATIONS READS

9 4,805

2 authors, including:

Andrea Cioncolini
The University of Manchester
88 PUBLICATIONS   1,277 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Flow-induced vibration in rod bundles: and experimental and computational benchmark study View project

High heat flux two-phase flow cooling for plasma facing components, including corrosion and magnetic field effects View project

All content following this page was uploaded by Andrea Cioncolini on 11 October 2017.

The user has requested enhancement of the downloaded file.


John R. Thome and Andrea Cioncolini, 2015. Encyclopedia of Two-Phase Heat Transfer and Flow, Set 1:
Fundamentals and Methods, Volume 3: Flow Boiling in Macro and Microchannels, World Scientific
Publishing.

Chapter 6

Two-Phase Pressure Drop

Accurate predictions of the pressure drops are critical for an energy efficient
design of any two-phase flow system. Moreover, the two-phase pressure drop is
required as input for determining the wall shear stress, a key flow parameter in
turbulence modeling to quantify the fluid-structure interaction. In this chapter,
after defining the pressure drop, discussing its practical relevance and describing
the most frequently used measuring techniques, the leading available prediction
methods for conventional channels and microchannels are reviewed and
critically analyzed.

1. Definition and Practical Relevance

In fluid flow problems in general, the term pressure is used to denote the static
pressure, which is the actual pressure in the fluid and depends on the state of the
fluid and not on its motion. The pressure gradient is defined as the gradient of
the scalar pressure field, and it is therefore a vector that points in the direction in
which the pressure undergoes the greatest rate of increase and that has a
magnitude equal to the rate of pressure increase in that direction, i.e. equal to the
directional space derivative of the pressure in that direction.
In the study of channel two-phase flow, it is customary to adopt a one-
dimensional approximation for space variables, and therefore assume the
pressure and all other flow parameters of interest to be constant over any channel
cross section and varying only in the channel axial direction z, as schematically
shown in Fig. 1. In this approximation, since the pressure decreases along the
flow, the pressure gradient is aligned with the channel axis, points upstream and
has a magnitude equal to the rate of change of the pressure along the channel
axis:
dP
∇P = − uz (1)
dz
where ∇ is the del operator, P is pressure, z is the coordinate measured along the
channel axis in the direction of the flow and uz is the unit vector in this direction.
As normally done in the two-phase flow literature, the magnitude of the pressure
1
2 J.R. Thome & A. Cioncolini

gradient dP/dz will be referred to as pressure gradient in the following sections,


even though strictly speaking this is a slight abuse of notation, as this ratio is a
scalar and not a vector.

Flow ∇P z

uz

Figure 1. Schematic representation of the one-dimensional approximation for channel two-phase


flow.

The difference ΔP = (P1-P2) between the pressures P1 and P2 at two arbitrary


channel cross sections z1 and z2 is referred to as the pressure drop between the
two cross sections, and its prediction is the object of this chapter. The ratio ΔP/Δz
= (P1-P2)/(z1-z2) of the pressure drop ΔP to the distance between the channel
cross sections Δz represents a specific pressure drop per unit channel length and
is particularly relevant in applications. Clearly, in the limit as z2 → z1, this ratio
tends to the pressure gradient dP/dz. As such, if the two channel cross sections
are close enough, then the ratio ΔP/Δz provides an experimentally accessible
approximation of the pressure gradient dP/dz. In the following, the term
measured pressure gradient will be tacitly assumed to mean a specific pressure
drop per unit channel length ΔP/Δz, measured between two channel cross
sections close enough together to provide a good approximation of the local
pressure gradient dP/dz.
For any two phase flow system, the accurate prediction of the pressure drop is
critical for a sound design and energy efficient operation, as the pressure drop
determines the pumping power required to run the system. The accurate
prediction of the pressure drop is critical for the design of natural circulation two-
phase flow loops (thermosyphons), used for example to extract decay heat from
nuclear reactors after shut-down. In direct expansion evaporators, typically a
pressure drop equivalent to a loss of 1‒2 K in saturation temperature from inlet to
outlet is set as the design limit. As such, if an evaporator is inaccurately designed
with a too low prediction of the pressure drop, then the system efficiency will be
penalized from the larger than expected fall in saturation temperature and
pressure across the evaporator. On the other hand, if the predicted pressure drop
is too large, then fewer tubes of longer length could have been utilized to obtain a
Two-Phase Pressure Drop 3

more compact and efficient unit. Finally, the accurate reconstruction of the
pressure profile along the channel is typically required in the post-processing of
heat transfer data taken under evaporating and condensing flow conditions, and
this requires an accurate pressure drop prediction method.

2. Measuring Techniques

The instrument that is most frequently used for measuring the pressure drop is
the DP cell, a differential pressure transducer that consists of two oil filled
chambers separated by an elastic sensing element, as shown schematically in Fig.
2. Each chamber is sealed and includes an elastic membrane that allows the oil in
the chamber to be pressurized externally. The DP cell is connected to the channel
with two manometric lines, i.e. two slender tubes that contain stagnant fluid and
allow the transmission of the pressure from the channel to the oil in the chambers
of the instrument. The pressure drop causes the pressure in one of the chambers
to be higher: this triggers a deformation of the elastic sensing element that is
converted to an electric signal proportional to the pressure drop.

Flow z
P1 P2

P1 P2

Manometric Line DP cell


Elastic Element

Figure 2. Schematic representation of the DP cell for pressure drop measurement.

Ideally, the holes drilled in the channel to connect the manometric lines
should be as small as possible, in order to minimize the perturbation to the flow.
In microchannels, realizing such connecting holes becomes extremely
challenging. With saturated fluids flowing in adiabatic conditions, an alternative
4 J.R. Thome & A. Cioncolini

is to measure the external temperature of the channel, as done by Revellin and


Thome (2007), Consolini (2008) and Ong (2010). If the channel is thermally
insulated, then the tube and the fluid are in thermal equilibrium so that measuring
the tube temperature provides the fluid saturation temperature, which in turn
allows determining the local saturation pressure from the fluid vapor-pressure
curve. The great advantage of this technique is that it is non-invasive, while its
limits are that it is applicable to adiabatic saturated two-phase flows only and its
poor sensitivity at small pressure drops.

Table 1(a). Experimental pressure drop databank.


Reference Fluids d (mm) P (MPa)
Silvestri et al. (1963) H2O-Steam; 3.25 a;5.00;5.00 a; 0.60-8.4
H2O-Ar; 5.20; 6.30; 7.00 a;
H2O+Alcohol-Ar 8.20; 10.1; 25.0
Gaspari et al. (1964) H2O-Steam 4.90; 4.99; 5.00; 2.0-9.2
5.04; 5.08; 8.07;
9.18; 15.2
Adorni et al. (1963a) H2O-Ar 15.1 0.60-2.2
Casagrande et al. (1963) H2O+Alcohol-Ar; 15.1; 25.0 0.29-2.4
H2O-N2
Cravarolo et al. (1964) H2O-Ar; 15.1; 25.0 0.60-2.1
Alcohol-Ar
Adorni et al. (1963b) H2O-Steam 15.2; 24.9 5.0
Würtz (1978) H2O-Steam 9.00 a; 10.0; 20.0 3.0-9.0
Anderson and Mantzouranis (1960) H2O-Air 10.8 0.10-0.18
Gill et al. (1964) H2O-Air 31.8 0.10-0.14
Shearer and Nedderman (1965) H2O-Air 15.9 0.11-0.21
Hinkle (1967) H2O-Air 12.6 0.25-0.62
Shiba and Yamazaki (1967) H2O-Air 24.5 0.10-0.16
Hewitt and Owen (1987) H2O-Air 31.8 0.24
Silva Lima et al. (2009) R134a 13.6 0.34-0.57
Revellin and Thome (2007) R134a; R245fa 0.517; 0.803 0.16-0.88
Consolini (2008) R134a; R245fa; 0.517; 0.803 0.16-0.79
R236fa
Ong (2010) R134a; R245fa; 1.03; 2.20; 3.04 0.14-0.93
R236fa
Steiner (1987) R12 14.0 0.15-0.31
Shedd (2010) R410a 0.508; 1.19; 2.96 1.89-3.07
Tibiriça and Ribatski (2011) R245fa 2.32 0.17-0.29
Zhang and Webb (2001) R22; R134a 3.25; 6.20 1.01-1.94
Dutkowski (2009; 2010) H2O-Air 1.05; 1.35; 1.68; 0.10-0.28
1.94; 2.30
Sekoguchi et al. (1968) H2O-Air 26.0 0.10-0.13
Kanno et al. (2010) H2O-Air; 0.30; 0.50; 1.00 0.11-0.47
Fluorinert-Air
Two-Phase Pressure Drop 5

3. Experimental Studies

A selection of studies that provide data on the pressure drop in two-phase flow is
summarized in Table 1. Selected histograms that further describe these data are
shown in Fig. 3.

Table 1(b). Experimental pressure drop databank with additional information.


Reference G (kgm-2s-1) x (1) (2)
Silvestri et al. (1963) 280-4577 0.01-0.91 ↑ 1217
Gaspari et al. (1964) 497-3949 0.02-0.99 ↑ 1060
Adorni et al. (1963a) 312-3420 0.06-0.82 ↑ 129
Casagrande et al. (1963) 255-2880 0.06-0.79 ↑ 115
Cravarolo et al. (1964) 266-2880 0.04-0.81 ↑ 840
Adorni et al. (1963b) 1100-3800 0.01-0.68 ↑ 104
Würtz (1978) 400-3000 0.08-0.70 ↑ 131
Anderson and Mantzouranis (1960) 50-1391 0.01-0.70 ↑ 33
Gill et al. (1964) 24-589 0.09-0.96 ↑ 177
Shearer and Nedderman (1965) 40-471 0.05-0.98 ↑ 100
Hinkle (1967) 155-733 0.15-0.71 ↑ 129
Shiba and Yamazaki (1967) 177-3530 0.01-0.79 ↑ 50
Hewitt and Owen (1987) 302-459 0.01-0.35 ↑ 41
Silva Lima et al. (2009) 308-523 0.01-0.99 → 656
Revellin and Thome (2007) 203-2026 0.01-0.96 → 1899
Consolini (2008) 272-1901 0.01-0.95 → 259
Ong (2010) 195-1517 0.01-0.98 → 1067
Steiner (1987) 50-246 0.10-0.81 → 158
Shedd (2010) 200-800 0.02-0.90 → 348
Tibiriça and Ribatski (2011) 100-701 0.10-0.98 → 38
Zhang and Webb (2001) 400-1000 0.20-0.85 → 38
Dutkowski (2009; 2010) 205-4000 0.01-0.21 → 156
Sekoguchi et al. (1968) 267-2661 0.01-0.03 → 172
Kanno et al. (2010) 100-500 0.02-0.96 → 194
(1): Flow direction: ↑ = vertical upflow; → = horizontal
(2): Number of data points
a
Hydraulic diameter (4 Aflow Pwet-1, Aflow = flow area, Pwet = wetted perimeter)

The databank in Table 1 contains 9111 measurements of the pressure drop


collected from 26 different literature studies that cover adiabatic flows through
vertical and horizontal channels, 13 different gas-liquid and vapor-liquid
combinations (both single-component saturated fluids such as water-steam and
refrigerants R12, R22, R134a, R236fa, R245fa, R410a and two-component fluids
such as water-air, water-argon, water-nitrogen, water plus alcohol-argon, alcohol-
argon and Fluorinert-air), operating pressures in the range of 0.1‒9.2 MPa and
tube diameters from 0.30 mm to 31.8 mm. In particular, as can be seen in Table
1, two studies addressed also non-circular channels: Silvestri et al. (1963)
measured the pressure drop in three annuli (hydraulic diameters of 3.25 mm, 5.00
6 J.R. Thome & A. Cioncolini

mm and 7.00 mm) while Würtz (1978) tested one annulus with a hydraulic
diameter of 9.00 mm. Both studies addressed operating conditions of interest for
nuclear reactor cooling applications (water-steam; operating pressures: 3.0‒9.0
MPa; mass flux: 500‒4500 kgm-2s-1).

2500 3000

2000
Nr Data Pts

Nr Data Pts
2000
1500

1000
1000
500

0 0
0 2 4 6 8 10 250 300 350 400 450 500 550 600
Pressure [MPa] Temperature [K]

1500 800
Nr Data Pts

Nr Data Pts

600
1000

400
500
200

0
0 1000 2000 3000 4000 5000 0
0 0.2 0.4 0.6 0.8 1
Mass Flux [kgm-2s-1] Vapor Quality

4000 1000
Nr Data Pts

800
Nr Data Pts

3000
600
2000 400

1000 200

0
-3 -2 -1 0 1 2 3 4
0 -1
0 5 10 15 20 25 30 35
Tube Diameter [mm] Log (Pressure Gradient [kPam ])
10

Figure 3. Selected histograms describing the pressure drop databank in Table 1.

In order to get some clue regarding the relation that links the measured
pressure gradient with the principal flow parameters, a selection of data from
Table 1 is presented in Figs. 4 through 9, where the measured pressure gradient is
plotted versus vapor quality in terms of operating pressure (Figs. 4 and 5), in
terms of mass flux (Figs. 6 and 7) and as a function of tube diameter (Figs. 8 and
9). As can be seen in Figs. 4 through 9, the measured pressure gradient increases
with an increase of vapor quality, a decrease of operating pressure, an increase of
mass flux and a decrease of the tube diameter.
Two-Phase Pressure Drop 7

25
P=1.89 MPa
P=3.07 MPa
Pressure Gradient [kPam ]
-1

20

15

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Vapor Quality

Figure 4. Pressure gradient vs. vapor quality: effect of operating pressure (R410a data of Shedd
(2010); Mass flux: 600 kgm-2s-1; Tube diameter: 2.96 mm).

90
P=3.0 MPa
80 P=5.0 MPa
Pressure Gradient [kPam ]
-1

P=7.0 MPa
70

60

50

40

30

20

10
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Vapor Quality

Figure 5. Pressure gradient vs. vapor quality: effect of operating pressure (H2O data of Würtz
(1978); Mass flux: 2000 kgm-2s-1; Tube diameter: 10.0 mm).

These observed trends of the pressure gradient can be intuitively interpreted


as follows. An increase of either the vapor quality or the mass flux triggers an
acceleration of the flow. This yields a steeper velocity profile at the channel wall
and, consequently, higher pressure gradients. An increase of the operating
pressure is followed by a reduction of the gas (or vapor) to liquid density ratio.
This yields a deceleration of the flow, a milder velocity profile at the channel
wall and, consequently, lower pressure gradients. Finally, a reduction of the tube
diameter while keeping a constant total mass flow rate yields an increase of the
mass flux, which means a steeper velocity profile at the channel wall and,
consequently, higher pressure gradients.
8 J.R. Thome & A. Cioncolini

It is worth highlighting that the above interpretation is no more than intuitive.


As a matter of fact, two-phase flows are so complicated that there is, as yet, no
convincing physical explanation of the observed trends of the pressure gradient.
Intuitively, the pressure gradient in two-phase flow should depend on the
turbulence structure within the flow, as happens with single-phase flows.

8 G=200 kgm-2s-1
G=400 kgm-2s-1
Pressure Gradient [kPam ]
-1

7
G=600 kgm-2s-1
6

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Vapor Quality

Figure 6. Pressure gradient vs. vapor quality: effect of mass flux (R410a data of Shedd (2010);
Pressure: 3.1 MPa; Tube diameter: 2.96 mm).

70
G=1000 kgm-2s-1
60 G=2000 kgm-2s-1
Pressure Gradient [kPam ]
-1

G=3000 kgm-2s-1
50

40

30

20

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Vapor Quality

Figure 7. Pressure gradient vs. vapor quality: effect of mass flux (H2O data of Würtz (1978);
Pressure: 7.0 MPa; Tube diameter: 10.0 mm).

Unfortunately, turbulence in channel two-phase flow is strongly flow pattern


dependent and is still largely unexplored. Besides, as noted by Brennen (2005),
two-phase flows are normally fluctuating and unsteady, and this yields a
multitude of quadratic interaction terms that contribute to the mean flow
Two-Phase Pressure Drop 9

similarly to Reynolds stress terms in turbulent single-phase flow, dramatically


complicating the physics of two-phase flows.

140
d=2.96 mm
120 d=1.19 mm
Pressure Gradient [kPam ]
-1

d=0.508 mm
100

80

60

40

20

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Vapor Quality

Figure 8. Pressure gradient vs. vapor quality: effect of tube diameter (R410a data of Shedd (2010);
Pressure: 1.9 MPa; Mass flux: 600 kgm-2s-1).

22
d=10 mm
20 d=20 mm
Pressure Gradient [kPam ]
-1

18

16

14

12

10

4
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Vapor Quality

Figure 9. Pressure gradient vs. vapor quality: effect of tube diameter (H2O data of Würtz (1978);
Pressure: 7.0 MPa; Mass flux: 1000 kgm-2s-1).

4. Theory and Prediction Methods

The theoretical starting point for predicting the pressure drop in channel two-
phase flow is the conservation equations for the total mass and total linear
momentum written in one-dimensional approximation. With reference to the one-
dimensional infinitesimal control volume depicted in Fig. 10, the total mass and
total linear momentum conservation equations are:
10 J.R. Thome & A. Cioncolini

d
( ρA) dz + Γl ( z + dz ) + Γg ( z + dz ) − Γl ( z ) − Γg ( z ) = 0 (2)
dt
d
(GA) dz + Γl ( z + dz ) Vl ( z + dz ) + Γg ( z + dz ) Vg ( z + dz ) −
dt
Γl ( z ) Vl ( z ) − Γg ( z ) Vg ( z ) = P( z ) A( z ) − P( z + dz ) A( z + dz ) (3)
− τ w ( z ) Pwet ( z ) dz − ρ ( z ) A ( z ) g sin(ϑ ) dz
where ρ is the cross sectional average density, defined in terms of the liquid and
gas (or vapor) densities ρl and ρg and cross sectional average void fraction ε as:
ρ = (1 − ε ) ρl + ε ρ g (4)

Control Volume

Flow z

z z+dz

Figure 10. One-dimensional infinitesimal control volume for mass and linear momentum
conservation equations.

The mass flow rates Γl and Γg and the average one-dimensional velocities Vl
and Vg of the liquid and gas (or vapor) phases in the channel are related as
follows:
(1 − x) G
Γl = (1 − x) Γ = ρl Vl Al = ρl Vl (1 − ε ) A ⇒ Vl = (5)
ρl (1 − ε )
xG
Γg = x Γ = ρ g Vg Ag = ρ g Vg ε A ⇒ Vg = (6)
ρg ε
where x is the vapor quality, Γ is the total mass flow rate flowing in the channel,
A is the cross sectional area of the channel, Al and Ag are the liquid and gas (or
vapor) flow areas (where Al + Ag = A), P is the pressure, τw is the wall shear stress
Two-Phase Pressure Drop 11

and Pwet is the channel wetted perimeter. Finally, g is the acceleration of gravity
and ϑ is the channel inclination with respect to the horizontal (ϑ = 0 for
horizontal flows, and positive values of the angle ϑ indicate upward flow).
In Equation (2), representing the conservation of the total mass, the first term
on the left hand side accounts for the accumulation of mass inside the control
volume, while the other terms represent the efflux of liquid and gas (or vapor) at
the two cross sections delimiting the control volume. In Equation (3),
representing the conservation of the total linear momentum, the first term on the
left hand side accounts for the accumulation of linear momentum inside the
control volume, while the other terms on the left hand side represent the efflux of
linear momentum at the two cross sections delimiting the control volume. The
terms on the right hand side in Eq. (3) represent the forces acting on the fluid
inside the control volume, taken with a positive sign if they assist the flow and
with a negative sign if they oppose the flow. The first two terms, in particular,
represent the pressure forces, the third term represents the fluid-channel frictional
interaction and the last term is the gravitational force acting on the fluid in the
control volume.
The functions in Eqs. (2) and (3) evaluated at (z + dz) can be expanded in
Taylor series, assuming they are suitably continuous and differentiable.
Neglecting higher order terms and rearranging yields:
∂ ∂
( ρA) + (GA) = 0 (7)
∂t ∂z
∂ ∂ ⎛ G2 A ⎞ ∂
(GA) + ⎜⎜ ⎟⎟ = − ( PA) − τ w Pwet − ρ A g sin(ϑ ) (8)
∂t ∂z ⎝ ρm ⎠ ∂z
where G is the mass flux, defined as the ratio of the total mass flow rate Γ to the
channel cross sectional area A:
Γ
G= (9)
A
while ρm is the momentum density:
−1
⎪ (1 − x) 2
⎧ x2 ⎫ ⎪
ρm = ⎨ + ⎬ (10)
⎩ ρl (1 − ε ) ρ g ε ⎪
⎪ ⎭
As can be seen, Eqs. (7) and (8) are two first-order, partial differential
equations that express the conservation of total mass and total linear momentum
in local form and in one-dimensional approximation. For steady-state flow
12 J.R. Thome & A. Cioncolini

conditions and for a channel of constant cross section these two conservation
equations become:
d
A (G ) = 0 ⇒ G = const (11)
dz
d ⎛ 1 ⎞ dP 4τ w
G2 ⎜⎜ ⎟⎟ = − − − ρ g sin(ϑ ) (12)
dz ⎝ ρm ⎠ dz dhyd
where dhyd is the channel hydraulic diameter, defined as four times the ratio of the
cross sectional flow area A to the wetted perimeter Pwet:
4A
d hyd = (13)
Pwet
Clearly, if the channel is a circular tube then the hydraulic diameter is equal to
the tube diameter. Conservation of total mass, therefore, requires the mass flux
(and total mass flow rate) be constant along the channel. The conservation
equation for total linear momentum, on the other hand, simplifies to a first order,
separable ordinary differential equation that can be recast in terms of the pressure
gradient as follows:

dP 4τ d ⎛ 1 ⎞ ⎛ dP ⎞ ⎛ dP ⎞ ⎛ dP ⎞
= − w − ρ g sin(ϑ ) − G 2 ⎜⎜ ⎟⎟ = ⎜ ⎟ + ⎜ ⎟ + ⎜ ⎟ (14)
dz d hyd dz ⎝ ρm ⎠ ⎝ dz ⎠ fr ⎝ dz ⎠ gr ⎝ dz ⎠acc
As can be seen, the pressure gradient on the left hand side of Eq. (14) is given
as the sum of three components. The first term is the frictional component of the
pressure gradient, and accounts for the fluid-structure interaction between the
two-phase flow and the bounding channel wall. This term is always negative for
cocurrent two-phase flows and represents pressure energy degraded due to
viscous dissipation within the flow. The second term is the gravitational
component of the pressure gradient, and accounts for the movement of the two-
phase flow within the earth gravitational force field. It is negative if the flow is
upwards, indicating that pressure energy is converted into potential gravitational
energy, while it is positive for downward flows, when potential gravitational
energy is converted into pressure energy. Finally, the third term is the
acceleration component of the pressure gradient, and accounts for the
acceleration or deceleration of the two-phase flow along the channel. This term
can be rearranged as follows:

d ⎛ 1 ⎞ d ⎛ G2 ⎞ d ⎛ G2 ⎞ d
G 2
⎜⎜ ⎟⎟ = ⎜⎜ ⎟⎟ = ⎜⎜ ρm 2 ⎟⎟ =
dz ⎝ ρm ⎠ dz ⎝ ρm ⎠ dz ⎝ ρm ⎠ dz
ρmVm2 ( ) (15)
Two-Phase Pressure Drop 13

G
Vm = = (1 − x)Vl + xVg (16)
ρm
As can be seen, the acceleration component represents the variation along the
channel of a specific kinetic energy for the two-phase flow, calculated with the
momentum density ρm and a vapor quality based average velocity Vm. This term
depends on the variations along the channel of the densities of the phases, of the
vapor quality and of the void fraction, as these parameters affect the momentum
density in Eq. (10). Typically, the acceleration term is negative in evaporating
flow conditions, as the flow is accelerating and pressure energy is therefore
converted into kinetic energy, while it is conversely positive with condensing
flows, when kinetic energy is converted into pressure energy as the flow
decelerates along the channel. In adiabatic flow conditions, the pressure
reduction along the channel triggers a decrease of the gas density, so that the
flow accelerates and the acceleration component of the pressure gradient is
therefore negative.
In practical applications, the pressure drop is predicted by integrating Eq. (14)
stepwise along the channel. As can be seen, this requires two closure models for
predicting the wall shear stress τw and the void fraction ε along the channel.
Besides, the liquid and gas densities ρl and ρg and the vapor quality x along the
channel are as well needed to carry out the integration. The most frequently used
pressure drop prediction methods are described below.

4.1 Homogeneous model

In the homogeneous model, the liquid and gas phases are assumed to flow in the
channel at the same velocity. The homogeneous void fraction εh is
correspondingly given as (see Chapter 4):
−1
⎛ 1 − x ρg ⎞ x
ε h = ⎜⎜1 + ⎟⎟ = (17)
⎝ x ρl ⎠ x + (1 − x) ρ g ρl−1
Accordingly, the cross sectional average density ρ and the momentum density
ρm become equal and coincide with the homogeneous density ρh:
−1
⎛1− x x ⎞
ρ = ρ m = ρ h = ⎜⎜ + ⎟

(18)
ρ
⎝ l ρ g ⎠

The pressure gradient is therefore as follows:


14 J.R. Thome & A. Cioncolini

dP G2 d ⎛ 1 ⎞
= −2 f h − ρh g sin(ϑ ) − G 2 ⎜⎜ ⎟⎟ (19)
dz ρh d hyd dz ⎝ ρ h ⎠
The homogeneous Fanning friction factor fh is typically calculated with
relations borrowed from single-phase smooth-pipe flow theory. Here, in
particular, the equations of Hagen‒Poiseuille and Filonenko are suggested, for
laminar and turbulent flow respectively:
−1
f h = 16 Re h for Re h < 2300
(20)
−1
f h = (1.58 ln Re h − 3.28) for Re h > 2300
where the homogeneous Reynolds number Reh is:
G d hyd
Re h = (21)
µh
A selection of available models for predicting the homogeneous viscosity µh
is summarized in Table 2, where µl and µg are the liquid and gas dynamic
viscosities while νl and νg are the liquid and gas kinematic viscosities.
Asymptotic consistency with single-phase all-liquid and all-gas flows requires
that µh → µl for x → 0+ and µh → µg for x → 1-.
The homogeneous viscosity definition proposed by McAdams et al. (1942)
was developed in formal analogy with the expression for the homogeneous
density, Eq. (18). The expression of Cicchitti et al. (1960), on the other hand, was
proposed on the basis of a better fit to available experimental data. The
expression of Dukler et al. (1964) is formally analogous to the one proposed by
Cicchitti et al. (1960) but is based on the kinematic viscosity. It is worth noting
that the homogeneous viscosity expressions of McAdams et al. (1942), Cicchitti
et al. (1960) and Dukler et al. (1964) are all asymptotically consistent with
single-phase all-liquid and all-gas flows.
Owens (1961) proposed probably the simplest available expression for the
homogeneous viscosity, setting it simply equal to the liquid viscosity. The author
argued that viscous dissipation should be mostly localized close to the channel
wall, as happens with single-phase flows. Since in most two-phase flow patterns
the liquid phase is in contact with the channel wall, the homogeneous viscosity
was then set equal to the liquid viscosity.

Table 2. Homogeneous two-phase flow viscosity definitions.


Authors Expression
Two-Phase Pressure Drop 15

−1
⎛1− x x ⎞
McAdams et al. (1942) µh = ⎜⎜ + ⎟
⎝ µl µ g ⎟⎠

Cicchitti et al. (1960) µ h = (1 − x )µl + xµ g

µh
Dukler et al. (1964) νh = = (1 − x )ν l + xν g
ρh

Owens (1961) µ h = µl

Beattie and Whalley (1982) µ h = µl (1 − ε h )(1 + 2.5ε h ) + ε h µ g

⎛ ρl − ρ g ⎞
Davidson et al. (1943) µh = µl ⎜⎜1 + x ⎟
⎝ ρ g ⎟⎠

ρh
Garcia et al. (2003) µ h = µl
ρl

2 µ l + µ g − 2( µ l − µ g ) x
Awad and Muzychka No. 1 (2008) µ h = µl
2 µl + µ g + ( µl − µ g ) x

2 µ g + µl − 2( µ g − µl )(1 − x)
Awad and Muzychka No. 2 (2008) µh = µ g
2µ g + µl + ( µ g − µl )(1 − x)

(3x − 1) µ g + (2 − 3x) µl +
1 ⎧⎪ ⎫⎪
Awad and Muzychka No. 3 (2008) µh = ⎨
4⎪
⎩ ((3x − 1)µ g + (2 − 3x)µl )2 + 8µl µ g ⎬⎪⎭
⎡ 2µl + µ g − 2( µl − µ g ) x ⎤
⎢ µl 2µ + µ + ( µ − µ ) x + ⎥
1 l g l g
Awad and Muzychka No. 4 (2008) µh = ⎢ ⎥
2 ⎢ 2µ g + µl − 2( µ g − µl )(1 − x) ⎥
⎢µ g ⎥
⎢⎣ 2µ g + µl + ( µ g − µl )(1 − x) ⎥⎦

The expression proposed by Beattie and Whalley (1982) was specifically


developed for bubbly flow and annular flow, while the expressions proposed by
Davidson et al. (1943) and Garcia et al. (2003) were proposed on the basis of a
better fit to available experimental data.
16 J.R. Thome & A. Cioncolini

As can be seen, the homogeneous viscosity expressions of Owens (1961),


Beattie and Whalley (1982), Davidson et al. (1943) and Garcia et al. (2003) are
all asymptotically consistent with single-phase all-liquid flow but not with single-
phase all-gas flow.
Finally, the four Awad and Muzychka (2008) homogeneous viscosity
expressions have been developed using a one-dimensional transport analogy
between thermal conductivity in porous materials and homogeneous viscosity in
two-phase flows. In their approach, the homogeneous viscosity is generated by
analogy to the effective thermal conductivity of a porous material. The first two
definitions from Awad and Muzychka (2008) in Table 2 are generated using the
Maxwell‒Eucken models for porous materials. Since these models assume one of
the phases to be continuous and the other dispersed, according to the authors the
corresponding homogeneous viscosity definitions should apply best to bubbly
flow and mist flow, where a continuous and a dispersed phase are clearly
recognizable. The third definition from Awad and Muzychka (2008) in Table 2 is
generated using the Effective Medium Theory for porous materials, where the
two phases are assumed to be randomly distributed in space, with neither one
being necessarily continuous or dispersed. Given the unsteady distribution of
phases in two-phase flow systems, according to them this homogeneous viscosity
definition should be applicable to any flow regime. Finally, the fourth
homogeneous viscosity definition from Awad and Muzychka (2008) in Table 2 is
just an average of the first two and is proposed as a simpler alternative to the
third definition. It is worth noting that the homogeneous viscosity expressions of
Awad and Muzychka (2008) are all asymptotically consistent with single-phase
all-liquid and all-gas flows.
Since the homogeneous model is based on neglecting the slip between the
phases, in principle it should be most effective for two-phase flows characterized
by a fine mixing of the phases, such as bubbly flow and mist flow, or in flows
with slight difference in density, such as saturated two-phase flows approaching
the critical state. On the other hand, the homogeneous model should be less
effective whenever the phases tend to become segregated and consequently
develop a slip, as happens with annular flow. As a matter of fact, however, the
homogeneous model can often handle separated flows quite well, provided that
the two-phase viscosity is properly defined.

4.2 Martinelli correlations

Martinelli and co-workers (Lockhart and Martinelli, 1949; Martinelli and Nelson,
1948) proposed the first correlations for the pressure drop in two-phase flow
Two-Phase Pressure Drop 17

specifically taking into account the fact that the two phases can have different
properties and, most importantly, different average velocities.
Lockhart and Martinelli (1949) provided graphical and tabular correlations for
the frictional pressure gradient and for the void fraction, expressed as functions
of the so-called Lockhart‒Martinelli parameter X defined as the square root of the
ratio of the single-phase frictional pressure gradients of the liquid and gas phases
flowing alone in the channel:
0.5 − 0.5 0.5 0.5
⎛ dP ⎞ ⎛ dP ⎞ ⎛ f ⎞ ⎛ 1 − x ⎞ ⎛ ρg ⎞
X = ⎜ ⎟ ⎜ ⎟ = ⎜ lo ⎟ ⎜ ⎟⎜ ⎟ (22)
⎝ dz ⎠lo ⎝ dz ⎠ go ⎜⎝ f go ⎟⎠ ⎝ x ⎠ ⎜⎝ ρl ⎟⎠
This parameter can be interpreted as a measure of how much the two-phase
flow behaves as a liquid rather than as a gas. The single-phase frictional pressure
gradients for the liquid-only and gas-only flows are calculated as follows:

⎛ dP ⎞ G 2 (1 − x) 2 ⎛ dP ⎞ G 2 x2
⎜ ⎟ = 2 f lo ; ⎜ ⎟ = 2 f go (23)
⎝ dz ⎠lo ρl d hyd ⎝ dz ⎠ go ρ g d hyd
where the single-phase liquid-only and gas-only Fanning friction factors for the
liquid flo and the gas fgo are calculated with the relations of Hagen‒Poiseuille and
McAdams (1954):

f = 16 Re −1 for Re < 1500


(24)
f = 0.046 Re − 0.2 for Re > 1500
Lockhart and Martinelli (1949) originally assumed laminar flow for Re ≤
1000, turbulent flow for Re ≥ 2000 and did not provide any clear indication for
the transition region 1000 < Re < 2000. The threshold of Re = 1500 proposed in
Eq. (24) simplifies the calculation and provides a smooth transition between the
Hagen‒Poiseuille and McAdams (1954) friction factor relations, as the values
calculated with these two expressions match at Re = 1500. Another possibility is
to assume turbulent flow in the transition region and use the McAdams friction
factor relation for 1000 < Re < 2000, as suggested by Stephan (1992). This
provides conservative pressure drop predictions, but introduces a discontinuity at
Re = 1000 that can yield convergence difficulties in numerical calculations. The
single-phase Reynolds numbers for the liquid-only Relo and the gas-only Rego
flows for use with Eq. (24) are defined as:
G (1 − x) d hyd G x d hyd
Relo = ; Re go = (25)
µl µg
18 J.R. Thome & A. Cioncolini

Notably, substituting all these expressions for both phases turbulent into Eq.
(22) yields:
0.9 0.5 0.1
⎛1− x ⎞ ⎛ ρg ⎞ ⎛ µl ⎞
X = X tt = ⎜ ⎟ ⎜⎜ ⎟⎟ ⎜ ⎟ ; for Relo , Re go > 1500 (26)
⎝ x ⎠ ⎜µ ⎟
⎝ ρl ⎠ ⎝ g ⎠

Table 3(a). Lockhart and Martinelli (1949) pressure drop correlation.


Liquid-only: Turbulent Liquid-only: Laminar
Gas-only: Turbulent Gas-only: Turbulent
X Φlo Φgo Φlo Φgo
0.01 128 1.28 120 1.20
0.02 68.4 1.37 64.0 1.28
0.04 38.5 1.54 34.0 1.36
0.07 24.4 1.71 20.7 1.45
0.1 18.5 1.85 15.2 1.52
0.2 11.2 2.23 8.90 1.78
0.4 7.05 2.83 5.62 2.25
0.7 5.04 3.53 4.07 2.85
1 4.20 4.20 3.48 3.48
2 3.10 6.20 2.62 5.25
4 2.38 9.50 2.05 8.20
7 1.96 13.7 1.73 12.1
10 1.75 17.5 1.59 15.9
20 1.48 29.5 1.40 28.0
40 1.29 51.5 1.25 50.0
70 1.17 82.0 1.17 82.0
100 1.11 111 1.11 111

The two-phase flow frictional pressure gradient is calculated as a multiple of


the single-phase pressure gradients defined in Eq. (23) as follows:

⎛ dP ⎞ ⎛ dP ⎞ 2 ⎛ dP ⎞ 2
⎜ ⎟ = ⎜ ⎟ Φlo = ⎜ ⎟ Φ go (27)
⎝ dz ⎠ fr ⎝ dz ⎠lo ⎝ dz ⎠ go
The correlations for the two-phase liquid-only and gas-only multipliers Φlo
and Φgo (note: not Φ2) are reported in tabular form in Table 3 and are reproduced
in Figs. 11 and 12, together with the following simple fitting equations proposed
by Chisholm (1967):
0.5 0.5
(
Φlo = 1 + CX −1 + X − 2 ) (
; Φ go = 1 + CX + X 2 ) (28)

Table 3(b). Lockhart and Martinelli (1949) pressure drop correlation.


Liquid-only: Turbulent Liquid-only: Laminar
Gas-only: Laminar Gas-only: Laminar
X Φlo Φgo Φlo Φgo
Two-Phase Pressure Drop 19

0.01 112 1.12 105 1.05


0.02 58.0 1.16 53.5 1.07
0.04 31.0 1.24 28.0 1.12
0.07 19.3 1.35 17.0 1.19
0.1 14.5 1.45 12.4 1.24
0.2 8.70 1.74 7.00 1.40
0.4 5.50 2.20 4.25 1.70
0.7 4.07 2.85 3.08 2.16
1 3.48 3.48 2.61 2.61
2 2.62 5.24 2.06 4.12
4 2.15 8.60 1.76 7.00
7 1.83 12.8 1.60 11.2
10 1.66 16.6 1.50 15.0
20 1.44 28.8 1.36 27.3
40 1.25 50.0 1.25 50.0
70 1.17 82.0 1.17 82.0
100 1.11 111 1.11 111

2
Turbulent-Turbulent
10 Laminar-Turbulent
Turbulent-Laminar
Two-Phase Multiplier

Laminar-Laminar
Liquid-Only

1
10

0
10 -2 -1 0 1 2
10 10 10 10 10
Lockhart-Martinelli Parameter

Figure 11. Lockhart and Martinelli (1949) pressure drop correlation: liquid-only two-phase
multiplier Φlo plotted versus the Lockhart‒Martinelli parameter X (Data points from Table 3 and
fitting equation, Eq. (28)).

As can be seen, four different correlation lines are defined depending on


whether the liquid-only and gas-only single-phase flows are laminar (Re < 1500)
or turbulent (Re > 1500). In particular, the laminar-laminar line is the lowest, the
turbulent-turbulent is the highest while the laminar-turbulent and turbulent-
laminar lines are intermediate.
20 J.R. Thome & A. Cioncolini

2 Turbulent-Turbulent
10 Laminar-Turbulent
Turbulent-Laminar
Two-Phase Multiplier

Laminar-Laminar
Gas-Only

1
10

0
10 -2 -1 0 1 2
10 10 10 10 10
Lockhart-Martinelli Parameter

Figure 12. Lockhart and Martinelli (1949) pressure drop correlation: gas-only two-phase
multiplier Φgo plotted versus the Lockhart‒Martinelli parameter X (Data points from Table 3 and
fitting equation, Eq. (28)).

The constant C in Eq. (28) is as indicated in Table 4:

Table 4. Values of the constant C in Eq. (28)


Liquid-Only Flow Gas-only Flow C
Turbulent Turbulent 20
Laminar Turbulent 12
Turbulent Laminar 10
Laminar Laminar 5

The correlation for the void fraction is reported in tabular form in Table 5 and
is reproduced in Fig. 13, together with the following fitting equation proposed by
Butterworth (1975):

ε = (1 + 0.28 X 0.71 )−1 (29)

Table 5. Lockhart and Martinelli (1949) void fraction correlation.


X 100 70 40 20 10 7 4
ε 0.1 0.16 0.24 0.34 0.47 0.52 0.60
X 2 1 0.7 0.4 0.2 0.1 0.07
ε 0.69 0.77 0.81 0.86 0.91 0.95 0.96

The experimental databank used by the authors covers two-phase flows of air
and different liquids (water, benzene, kerosene and different oils), operating
pressures in the range of 0.1‒0.4 MPa and tube diameters from 1.5 mm to 25.8
mm. As can be seen in Tables 3 and 5, this method is limited to Lockhart‒
Two-Phase Pressure Drop 21

Martinelli parameter values in the range of 10-2‒102, although the fitting


equations Eqs. (28) and (29) can be extrapolated beyond these limits. As noted
by Hewitt and Hall-Taylor (1970), the theoretical justification of this correlation
included in the original publication is at best obscure, so that the method should
be regarded as purely empirical. Notwithstanding its limitations and the
availability of more modern and accurate correlations, the Lockhart and
Martinelli correlation is still actively used, either directly and as a basis for ‘add
on’ methods where the constant C in Eq. (28) becomes a correlated parameter, as
will be seen later.

0
10
Void Fraction

Data Points from Table 5


Fitting Equation, Eq. (29)
-1
10 -2 -1 0 1 2
10 10 10 10 10
Lockhart-Martinelli Parameter

Figure 13. Lockhart and Martinelli (1949) void fraction correlation.

An upgraded version of this method, specifically designed for high pressure


steam-water boiling flows, was proposed by Martinelli and Nelson (1948). The
two-phase flow frictional pressure gradient is now calculated as a multiple of the
single-phase liquid-all pressure gradient as follows:

⎛ dP ⎞ ⎛ dP ⎞ 2
⎜ ⎟ = ⎜ ⎟ Φla (30)
⎝ dz ⎠ fr ⎝ dz ⎠la
where the liquid-all single-phase frictional pressure gradient is that which would
result if the entire mass flow rate flowed as liquid through the channel, calculated
using the Blasius (1913) single-phase equation:

⎛ dP ⎞ G 2
− 0.25 G d hyd
⎜ ⎟ = 2 fla ; fla = 0.0791 Rela ; Rela = (31)
⎝ dz ⎠la ρl d hyd µl
The two-phase multiplier Φla2 is provided in tabular form as function of the
vapor quality and operating pressure in Table 6. It is worth highlighting that the
22 J.R. Thome & A. Cioncolini

Martinelli and Nelson (1948) correlation is only partially supported by data, as


the curves used to derive the two-phase multiplier values included in Table 6
were established by trial and error using the limited water-steam data available as
a guide.

Table 6(a). Martinelli and Nelson (1948) two-phase multiplier Φla2


Operating Pressure [MPa]
x 0.101 0.689 3.44 6.89 10.3
0.01 5.6 3.5 1.8 1.6 1.35
0.05 30 15 5.3 3.6 2.4
0.10 69 28 8.9 5.4 3.4
0.20 150 56 16.2 8.6 5.1
0.30 245 83 23.0 11.6 6.8
0.40 350 115 29.2 14.4 8.4
0.50 450 145 34.9 17.0 9.9
0.60 545 174 40.0 19.4 11.1
0.70 625 199 44.6 21.4 12.1
0.80 685 216 48.6 22.9 12.8
0.90 720 210 48.0 22.3 13.0
1.00 525 130 30.0 15.0 8.6

Table 6(b). Martinelli and Nelson (1948) two-phase multiplier Φla2


Operating Pressure [MPa]
x 13.8 17.2 20.7 22.1
0.01 1.2 1.1 1.05 1.00
0.05 1.75 1.43 1.17 1.00
0.10 2.45 1.75 1.30 1.00
0.20 3.25 2.19 1.51 1.00
0.30 4.04 2.62 1.68 1.00
0.40 4.82 3.02 1.83 1.00
0.50 5.59 3.38 1.97 1.00
0.60 6.34 3.70 2.10 1.00
0.70 7.05 3.96 2.23 1.00
0.80 7.70 4.15 2.35 1.00
0.90 7.95 4.20 2.38 1.00
1.00 5.90 3.70 2.15 1.00

4.3 Baroczy‒Chisholm correlation

Baroczy (1966) proposed a graphical correlation to predict the frictional pressure


gradient, intended for application to a variety of gas-liquid combinations of
interest in practical applications. Successively, Chisholm (1973) developed the
empirical equations that approximated the graphs of Baroczy (1966), greatly
simplifying the practical application of this correlation. In this method, the two-
Two-Phase Pressure Drop 23

phase frictional pressure gradient is predicted as a multiple of the single-phase


liquid-all pressure gradient as follows:

⎛ dP ⎞ ⎛ dP ⎞ 2
⎜ ⎟ = ⎜ ⎟ Φ la (32)
⎝ dz ⎠ fr ⎝ dz ⎠la
where the two-phase multiplier Φla2 is calculated as:

(
Φ la2 = 1 + (Y 2 − 1) B x 0.875 (1 − x)0.875 + x1.75 ) (33)

The physical property coefficient Y 2 is defined as the square root of the ratio
of the single-phase frictional pressure gradients that would result if the entire
two-phase mass flow rate flowed through the channel as either liquid or gas:
0.5 −0.5
⎛ dP ⎞ ⎛ dP ⎞
Y =⎜ ⎟ ⎜ ⎟ (34)
⎝ dz ⎠la ⎝ dz ⎠ ga
This parameter is similar to the Lockhart‒Martinelli parameter X in Eq. (22)
and can be similarly interpreted as a measure of how much the two-phase flow
behaves as a liquid rather than as a gas. The advantage of the Chisholm
parameter Y is that it is independent of vapor quality and therefore does not
change along the channel, thus simplifying evaporation and condensation
calculations. The liquid-all and gas-all single-phase frictional pressure gradients
are calculated using the Blasius (1913) single-phase equation as follows:

⎛ dP ⎞ G 2
− 0.25 G d hyd
⎜ ⎟ = 2 f la ; f la = 0.0791Re la ; Re la = (35)
⎝ dz ⎠la ρl d hyd µl

⎛ dP ⎞ G 2
− 0.25 G d hyd
⎜ ⎟ = 2 f ga ; f ga = 0.0791Re ga ; Re ga = (36)
⎝ dz ⎠ ga ρ g d hyd µg
The dimensionless parameter B in Eq. (33) is calculated as a function of the
Chisholm parameter Y and total mass flux G (note: to be entered in kgm-2s-1) as
indicated in Table 7:

Table 7. Values of the parameter B in Eq. (33)


Y G [kgm-2s-1] B
≤ 9.5 < 500 4.8
500‒1900 2400 G -1
≥ 1900 55 G -0.5
9.5‒28 ≤ 600 520 (YG 0.5)-1
> 600 21 Y -1
≥ 28 All values 15000 (Y 2G 0.5)-1
24 J.R. Thome & A. Cioncolini

The correlation for the void fraction is provided by Baroczy (1966) in


graphical form and can be approximated with the following fitting equation
(Butterworth, 1975):
0.65 0.13 −1
⎛ 0.74 ⎞
⎛ 1 − x ⎞ ⎛⎜ ρ g ⎞ ⎛ µl ⎞
ε = ⎜⎜1 + ⎜ ⎟ ⎟⎟ ⎜ ⎟ ⎟ (37)
⎝ x ⎠ ⎜⎝ ρl ⎠
⎜µ
⎝ g



⎝ ⎠

4.4 Friedel correlation

In the method of Friedel (1979), the two-phase frictional pressure gradient is


predicted as a multiple of the single-phase liquid-all pressure gradient as follows:

⎛ dP ⎞ ⎛ dP ⎞ 2
⎜ ⎟ = ⎜ ⎟ Φ la (38)
⎝ dz ⎠ fr ⎝ dz ⎠la
where the liquid-all single-phase frictional pressure gradient is that which would
result if the entire mass flow rate flowed as liquid through the channel, calculated
using the Blasius (1913) single-phase equation:

⎛ dP ⎞ G 2
− 0.25 G d hyd
⎜ ⎟ = 2 f la ; f la = 0.0791Re la ; Re la = (39)
⎝ dz ⎠la ρl d hyd µl
while the two-phase multiplier Φla2 is correlated as:
3.24 A2 A3
Φ la2 = A1 + (40)
Fr 0.045 We0.035
The Froude number Fr and the Weber number We are:

G2 G 2 dhyd
Fr = ; We = (41)
ρh2 g dhyd ρh σ
where ρh is the homogeneous density defined in Eq. (18) and σ is the surface
tension. The parameters A1, A2 and A3 are defined as follows:

2
⎛ ρ ⎞⎛ f ⎞
A1 = (1 − x ) + x 2 ⎜ l ⎟⎜⎜ ga ⎟⎟ (42)
⎜ρ ⎟ f
⎝ g ⎠⎝ la ⎠
0.224
A2 = x 0.78 (1 − x ) (43)
Two-Phase Pressure Drop 25

0.91 0.19 0.7


⎛ρ ⎞ ⎛ µg ⎞ ⎛ µg ⎞
A3 = ⎜ l ⎟ ⎜⎜ ⎟⎟ ⎜⎜1 − ⎟⎟ (44)
⎜ρ ⎟
⎝ g⎠ ⎝ µl ⎠ ⎝ µl ⎠
where fga is the Fanning friction factor for the entire mass flow rate flowing as
gas and is calculated using the Blasius (1913) single-phase equation:

− 0.25 G d hyd
f ga = 0.0791Re ga ; Re ga = (45)
µg
The author does not provide a void fraction correlation to predict the
acceleration and gravitational components of the pressure gradient, so that the
user can pick the void fraction prediction method that best fits his/her application
(see Chapter 4). The experimental databank used by the author contained about
16000 data points for horizontal and vertical upflow conditions, 13 fluids (water-
steam, R11, R12, R22, R113, nitrogen, water-air, oil-air, water-methane, oil-
methane, water-nitrogen, alcohol-argon and water-argon), operating pressures in
the range of 0.06‒21 MPa, tube diameters from 0.98 mm to 257.4 mm and both
circular and non-circular channels (rectangular and annuli).

4.5 Müller-Steinhagen and Heck correlation

According to Müller-Steinhagen and Heck (1986), the two-phase frictional


pressure gradient is predicted as follows:

⎛ dP ⎞ ⎧⎪⎛ dP ⎞ ⎡⎛ dP ⎞ ⎛ dP ⎞ ⎤ ⎫⎪ 0.33 ⎛ dP ⎞ 3.00


⎜ ⎟ = ⎨⎜ ⎟ + 2 ⎢⎜ ⎟ − ⎜ ⎟ ⎥ x ⎬ (1 − x ) + ⎜ ⎟ x (46)
⎝ dz ⎠ fr ⎪⎩⎝ dz ⎠la ⎣⎢⎝ dz ⎠ ga ⎝ dz ⎠la ⎦⎥ ⎪⎭ ⎝ dz ⎠ ga
where the single-phase liquid-all and gas-all pressure gradients are:
2 2
⎛ dP ⎞ G ⎛ dP ⎞ G
⎜ ⎟ = 2 f la ; ⎜ ⎟ = 2 f ga (47)
⎝ dz ⎠la ρl d hyd ⎝ dz ⎠ ga ρ g d hyd
The single-phase Fanning friction factor fla and fga for the liquid-all and gas-
all flows are estimated with the relations of Hagen‒Poiseuille and Blasius (1913):

f = 16 Re −1 for Re < 1187


(48)
f = 0.0791Re − 0.25 for Re > 1187
using the liquid-all and gas-all Reynolds numbers defined as:
26 J.R. Thome & A. Cioncolini

G d hyd G d hyd
Rela = ; Re ga = (49)
µl µg
The value of the Reynolds number threshold between laminar and turbulent
flow adopted by the authors provides a smooth transition between the Hagen‒
Poiseuille and Blasius (1913) friction factor relations, as the values calculated
with these two expressions match at Re = 1187. The authors do not provide a
void fraction correlation to predict the acceleration and gravitational components
of the pressure gradient, so that the user can pick the void fraction prediction
method that best fits his/her application (see Chapter 4). The experimental
databank used by the authors contained 9313 data points for circular tubes in
horizontal, vertical upflow and vertical downflow conditions, 9 fluids (water-
steam, R11, R12, argon, nitrogen, neon, water-air, water-argon and oil-air),
unspecified operating pressure range and tube diameters from 4.00 mm to 392
mm.

4.6 Lombardi and Carsana correlation

In the method proposed by Lombardi and Carsana (1992), the two-phase


frictional pressure gradient is predicted as follows:
2
⎛ dP ⎞ G
⎜ ⎟ = 2 f tp (50)
⎝ dz ⎠ fr ρ h d hyd
where ρh is the homogeneous density defined in Eq. (18) while ftp is a two-phase
Fanning friction factor calculated as a blend of a single-phase liquid-all fla, a
single-phase gas-all fga and a mixture fm Fanning friction factors as follows:
f tp = f la bl + f ga bg + f m bm (51)

The single-phase Fanning friction factors for the liquid-all and gas-all single-
phase flows are calculated with the relations of Hagen‒Poiseuille and Selander
(1978):

f = 16 Re −1 for Re < 2400


−2
⎡ ⎛ 10 r ⎞⎤ (52)
f = ⎢3.8Log10 ⎜ + 0.2 ⎟⎥ for Re > 2400
⎢⎣ ⎜ Re d hyd ⎟⎥
⎝ ⎠⎦
where r is the channel surface roughness and Re the single-phase Reynolds
number. The mixture Fanning friction factor fm is calculated as:
Two-Phase Pressure Drop 27

f m = 1.38 Bo We−1.25 for We ≤ 30 Bo (53)

f m = 0.046 We−0.25 for We > 30 Bo (54)

Notably, at We = 30Bo the values of fm calculated with the two expressions


provided match. Bo and We are modified Bond and Weber numbers:
0.5
ρl g (dhyd − d0 )2 ⎛ µ g ⎞ G 2 dhyd ⎛ µg ⎞
Bo = ⎜ µ ⎟; We = ρ σ
⎜ ⎟ ⎜⎜ ⎟⎟ (55)
σ ⎝ l⎠ h ⎝ µl ⎠
The parameter d0 in the Bond number above is a reference tube diameter with
a value d0 = 1.0 mm. If the channel hydraulic diameter dhyd is smaller than d0,
then the Bond number Bo is reset to zero. Finally, the weight coefficients bl, bg
and bm in Eq. (51) are calculated as follows:
2 ρl / ρ g 600 ρ g / ρ l
bl = (1 − x ) ; bg = x ; bm = 1 − bl − bg (56)

The authors suggest the homogeneous void fraction in Eq. (17) be used to
predict the acceleration and gravitational components of the pressure gradient, so
that these two pressure gradient components are formally as indicated in Eq. (19).
The experimental databank used by the authors contained about 15000 data
points for vertical upflow conditions, both adiabatic and evaporating flow
conditions, 5 fluids (water-steam, water-nitrogen, water-argon, water plus
alcohol-argon and alcohol-argon), operating pressures in the range of 0.3‒9.2
MPa, tube diameters from 0.2 mm to 446 mm and both circular and non-circular
channels (annuli and rod bundles).

4.7 Shannak correlation

According to Shannak (2008), the two-phase frictional pressure gradient is


predicted as follows:
2
⎛ dP ⎞ G
⎜ ⎟ = 2 f tp (57)
⎝ dz ⎠ fr ρ h d hyd
where ρh is the homogeneous density defined in Eq. (18) while ftp is a two-phase
Fanning friction factor calculated using the single-phase turbulent flow equation
of Chen (1979) as follows:
28 J.R. Thome & A. Cioncolini

−2
⎡ ⎛ 2r 16.2426 ⎞⎤
ftp = ⎢3.48 − 1.7372 ln⎜ − ln A ⎟⎥ (58)
⎢⎣ ⎜d Retp ⎟⎥
⎝ hyd ⎠⎦
1.1098 0.8981
⎛ 2r ⎞ ⎛ 7.149 ⎞
A = 0.164 ⎜ ⎟ +⎜ ⎟ (59)
⎜d ⎟ ⎜ Re ⎟
⎝ hyd ⎠ ⎝ tp ⎠

where r is the channel surface roughness while Retp is the two-phase flow
Reynolds number defined as:

⎡ 2 ρ ⎤
G d hyd ⎢ x 2 + (1 − x ) g ⎥
⎣ ρl ⎦
Retp = (60)
ρ
µ g x + µl (1 − x ) g
ρl
The author does not provide a void fraction correlation to predict the
acceleration and gravitational components of the pressure gradient, so that the
user can pick the void fraction prediction method that best fits his/her application
(see Chapter 4). The experimental databank used by the author is the same one
used by Friedel (1979): about 16000 data points for horizontal and vertical
upflow conditions, 13 fluids (water-steam, R11, R12, R22, R113, nitrogen,
water-air, oil-air, water-methane, oil-methane, water-nitrogen, alcohol-argon and
water-argon), operating pressures in the range of 0.06‒21 MPa, tube diameters
from 0.98 mm to 257.4 mm and both circular and non-circular channels
(rectangular and annuli).

4.8 Cioncolini et al. annular flow correlation

Cioncolini et al. (2009) proposed a pressure drop correlation specifically


designed for annular two-phase flows that explicitly takes into account the
fraction of liquid that is entrained as droplets in the gas core flow. The two-phase
frictional pressure gradient is predicted as follows:

⎛ dP ⎞ ρ V2
⎜ ⎟ = 2 f tp c c (61)
⎝ dz ⎠ fr d hyd
where the two-phase Fanning friction factor ftp is:

ftp = 0.172 Wec−0.372 for Bo > 4 (macrochannels) (62)


Two-Phase Pressure Drop 29

ftp = 0.0196 Wec−0.372 Relf0.318 for Bo < 4 (microchannels) (63)

The core flow Weber number Wec, the liquid film Reynolds number Relf and
the Bond number Bo are as follows:
2
ρc Vc2 dc G d hyd g ( ρl − ρ g ) d hyd
Wec = ; Relf = (1 − e)(1 − x) ; Bo = (64)
σ µl σ
where e is the entrained liquid fraction, representing the ratio of the entrained
liquid droplets mass flow rate to the total liquid mass flow rate (see Chapter 5).
The droplet-laden core flow density ρc, average core velocity Vc and core
diameter dc are calculated neglecting the slip between the carrier gas phase and
the entrained liquid droplets as follows:

x + e (1 − x) xG xρl + e (1 − x) ρ g
ρc = ; Vc = ; d c = d hyd ε (65)
x e (1 − x) ε ρ g x ρl
+
ρg ρl
The momentum density in the acceleration component of the pressure
gradient generalizes as follows for annular flow with entrained liquid droplets:
−1
2 2
⎪ (1 − e) (1 − x) x e x (1 − x)
⎧ x2 ⎫ ⎪
ρm = ⎨ + + ⎬ (66)
⎩ (1 − ε ) x ρl
⎪ ε ρg ε ρ g ⎪⎭
As can be seen, this correlation requires two closure models for predicting the
entrained liquid fraction e and the void fraction ε. In the original formulation of
the correlation, the methods of Oliemans et al. (1986) (see Chapter 5) and
Woldesemayat and Ghajar (2007) (see Chapter 4) were used. Now these closure
laws can be replaced with the more accurate ones of Cioncolini and Thome
(2012a,b) (see Chapters 4‒5). The experimental databank used by the authors
contains 3908 data points for adiabatic annular flow, circular tubes, horizontal
and vertical upflow conditions, 8 fluids (water-steam, R134a, R245fa, water-air,
water-nitrogen, alcohol-argon, water plus alcohol-argon and water-argon),
operating pressures in the range of 0.1‒9.4 MPa and tube diameters from 0.517
mm to 31.7 mm.

4.9 Mishima correlations

Mishima and co-workers (Mishima and Hibiki, 1996; Sun and Mishima, 2009)
proposed two modifications of the Lockhart and Martinelli correlation
30 J.R. Thome & A. Cioncolini

specifically derived to better fit microscale data. The two-phase frictional


pressure gradient is predicted as follows:

⎛ dP ⎞ ⎛ dP ⎞ 2
⎜ ⎟ = ⎜ ⎟ Φ lo (67)
⎝ dz ⎠ fr ⎝ dz ⎠lo
where the liquid-only single-phase frictional pressure gradient that would result if
the liquid part of the total mass flow rate flowed alone through the channel is
calculated as:

⎛ dP ⎞ G 2 (1 − x) 2
⎜ ⎟ = 2 f lo (68)
⎝ dz ⎠lo ρl d hyd
The single-phase liquid-only Fanning friction factor flo is calculated with the
relations of Hagen‒Poiseuille and Blasius (1913):

flo = 16 Relo−1 for Relo < 2000


(69)
flo = 0.0791Relo− 0.25 for Relo > 2000
The single-phase Reynolds numbers for the liquid-only Relo is defined as:
G (1 − x) d hyd
Relo = (70)
µl
The liquid-only two-phase multiplier Φlo2 is:

Φ lo2 = 1 + CX −1 + X −2 (71)

According to Mishima and Hibiki (1996), the constant C appearing in Eq.


(71) can be treated as a correlating parameter, and it is calculated as follows:
C = 21 ( 1 − exp (−0.319 d hyd ) ) (72)

where the hydraulic diameter dhyd is to be entered in mm. The experimental


databank used to derive Eq. (72) covers horizontal and vertical upflow
conditions, circular and rectangular channels, 3 fluids (water-air, ammonia-steam
and R113-nitrogen) and channel diameters from 0.07 mm to 4.08 mm.
In the method of Sun and Mishima (2009), on the other hand, the constant C
appearing in Eq. (71) is predicted as:
0.1 0.4
⎛1− x ⎞ ⎛ µl ⎞
C = 1.79 X − 0.19 ⎜ ⎟ ⎜ ⎟ (73)
⎝ x ⎠ ⎜µ ⎟
⎝ g⎠
Two-Phase Pressure Drop 31

where X is the Lockhart-Martinelli parameter defined in Eq. (22). The


experimental databank used to derive Eq. (73) contains 2092 points and covers
horizontal and vertical upflow conditions, circular and non-circular channels
(rectangular, triangular, annulus), 11 fluids (water-air, R123, R134a, R22,
R236ea, R245fa, R404a, R407c, R410a, R507, CO2) and channel hydraulic
diameters from 0.506 mm to 12.0 mm.
Friedel (1979)
4
Log10(Pressure Gradient [kPam ]):
-1

2
Experimental

-1

-2
-2 -1 0 1 2 3 4
-1
Log10(Pressure Gradient [kPam ]): Predicted

Muller-Steinhagen and Heck (1986)


4
Log10(Pressure Gradient [kPam ]):
-1

2
Experimental

-1

-2
-2 -1 0 1 2 3 4
-1
Log10(Pressure Gradient [kPam ]): Predicted

Sun and Mishima (2009)


4
Log10(Pressure Gradient [kPam ]):
-1

2
Experimental

-1

-2
-2 -1 0 1 2 3 4
-1
Log10(Pressure Gradient [kPam ]): Predicted

Figure 14. Pressure gradient: experimental data of Table 1 vs. predictions of Friedel (1979)
correlation [top], predictions of Müller-Steinhagen and Heck (1986) correlation [middle] and
predictions of Sun and Mishima (2009) correlation [bottom].
32 J.R. Thome & A. Cioncolini

4.10 Comparison of above methods to experimental database

With respect to the databank in Table 1, the best prediction methods for the
pressure drop among those presented above are the correlations of Friedel (1979),
Müller-Steinhagen and Heck (1986) and Sun and Mishima (2009). Measured
data from Table 1 are compared with the predictions of these methods in Fig. 14.
Typical mean absolute percentage errors are on the order of 30 %. Based on
available data, these prediction methods can be extrapolated to non-circular
channels using the hydraulic diameter (4 Aflow Pwet-1) in place of the tube diameter.
For annular flows, the best predictions are given by the Cioncolini et al. (2009)
correlation.

5. Nomenclature

A, Aflow channel cross section flow area (m2)


Ag channel cross sectional area occupied by the gas phase (m2)
Al channel cross sectional area occupied by the liquid phase (m2)
b weight coefficient (-)
B Chisholm parameter (-)
Bo Bond number (-)
dhyd channel hydraulic diameter (m)
e entrained liquid fraction (-)
f Fanning friction factor (-)
Fr Froude number (-)
g acceleration of gravity (ms-2)
G mass flux (kgm-2s-1)
P pressure (Pa)
Pwet channel wetted perimeter (m)
r channel surface roughness (m)
Re Reynolds number (-)
Vg one-dimensional gas velocity (ms-1)
Vl one-dimensional liquid velocity (ms-1)
We Weber number (-)
x vapor quality (-)
X Martinelli number (-)
Y Chisholm parameter (-)
z length alone channel (m)
Two-Phase Pressure Drop 33

Γ total mass flow rate (kgs-1)


Γl liquid mass flow rate (kgs-1)
Γg gas mass flow rate (kgs-1)
µg gas viscosity (kgm-1s-1)
µl liquid viscosity (kgm-1s-1)
ε cross sectional void fraction (-)
ρg vapor density (kgm-3)
ρl liquid density (kgm-3)
σ surface tension (kgs-2)
τ shear stress (Nm-2)
Φ two-phase multiplier (-)

6. References

Adorni, N., Casagrande, I., Cravarolo, L., Hassid, A., Pedrocchi, E. and Silvestri, M., (1963a).
Further investigations in adiabatic dispersed flow: pressure drop and film thickness
measurements with different channel geometries-analysis of the influence of geometrical and
physical parameters. CISE Report R-53, Segrate, Italy.
Adorni, N., Alessandrini, A., Peterlongo, G. and Ravetta, R., (1963b). Large scale experiments on
heat transfer and hydrodynamics with steam-water mixtures. CISE Report R-79, Segrate,
Italy.
Anderson , G.H. and Mantzouranis, B.G., (1960). Two-phase (gas-liquid) flow phenomena-I:
pressure drop and hold-up for two-phase flow in vertical tubes. Chem. Eng. Sci. 12, pp. 109-
126.
Awad, M.M. and Muzychka, Y.S., (2008). Effective property models for homogeneous two-phase
flows. Exp. Therm. Fluid Sci. 33, pp. 106-113.
Baroczy, C.J., (1966). A systematic correlation for two-phase pressure drop. Chem. Eng. Prog.
Symp. Ser. 62, pp. 232-249.
Blasius, H., (1913). Das ähnlichkeitsgesets bei reibungsvorgängen in flüssigkeiten. Forschg. Arb.
Ing.-Wes., pp. 131.
Beattie, D.R.H. and Whalley, P.B, (1982). Simple two-phase frictional pressure drop calculation
method, Int. J. Multiphase Flow 8, pp. 83-87.
Brennen, C.E., (2006). Fundamentals of Multiphase Flow. Cambridge University Press, New York,
USA.
Butterworth, D., (1975). A comparison of some void fraction relationships for co-current gas-liquid
flow. Int. J. Multiphase Flow 1, pp. 845-850.
Casagrande, I., Cravarolo, L., Hassid, A. and Pedrocchi, E., (1963). Adiabatic dispersed two-phase
flow: further results on the influence of physical properties on pressure drop and film
thickness. CISE Report R-73, Segrate, Italy.
Chen, N.H., (1979). An explicit equation for friction factor in pipes. Ind. Eng. Chem. Fund. 18, pp.
296-297.
34 J.R. Thome & A. Cioncolini

Chisholm, D., (1967). A theoretical basis for the Lockhart-Martinelli correlation for two-phase
flow. Int. J. Heat Mass Transfer 10, pp. 1767-1778.
Chisholm, D., (1973). Pressure drop due to friction during the flow of evaporating two-phase
mixtures in smooth tubes and channels. Int. J. Heat Mass Transfer 16, pp. 347-358.
Cicchitti, A., Lombardi, C., Silvestri, M., Soldaini, G. and Zavattarelli, R., (1960). Two-phase
cooling experiments-pressure drop, heat transfer and burnout experiments. Energia Nucleare
7, pp. 407-425.
Cioncolini, A., Thome, J.R. and Lombardi, C., (2009). Unified macro-to-microscale method to
predict two-phase frictional pressure drops of annular flows. Int. J. Multiphase Flow 35, pp.
1138-1148.
Cioncolini, A. and Thome, J.R., (2012a). Void fraction prediction in annular two-phase flow. Int. J.
Multiphase Flow 43, pp. 72-84.
Cioncolini, A. and Thome, J.R., (2012b). Entrained liquid fraction prediction in adiabatic and
evaporating annular two-phase flow. Nucl. Eng. Des. 243, pp. 200-213.
Consolini, L., (2008). Convective boiling heat transfer in a single micro-channel. Ph.D. thesis,
Swiss Federal Institute of Technology-EPFL, Lausanne, Switzerland.
Cravarolo, L., Hassid, A. and Pedrocchi, E., (1964). Further investigation on two-phase adiabatic
annular-dispersed flow: effect of length and some inlet conditions on flow parameters. CISE
Report R-93, Segrate, Italy.
Davidson, W.F., Hardie, P.H., Humphreys, C.G.R., Markson, A.A., Mumford, A.R. and Ravese, T.,
(1943). Studies of heat transmission through boiler tubing at pressures from 500 to 3000 Lbs.
Trans. ASME 65, pp. 553-591.
Dukler, A.E., Wicks, M. and Cleveland, R.G., (1964). Frictional pressure drop in two-phase flow,
Part A: a comparison of existing correlations for pressure loss and holdup and Part B: an
approach through similarity analysis. AIChE J. 10, pp. 38-51.
Dutkowski, K., (2009). Two-phase pressure drop of air-water in minichannels. Int. J. Heat Mass
Transfer 52, pp. 5185-5192.
Dutkowski, K., (2010). Air-water two-phase frictional pressure drop in minichannels. Heat
Transfer Eng. 31, pp. 321-330.
Friedel, L., (1979). Improved friction pressure drop correlation for horizontal and vertical two-
phase pipe flow. European Two-Phase Flow Group Meeting, paper E2, Ispra, Italy.
Garcia, F., Garcia, R., Padrino, J.C., Mata, C., Trallero, J.L. and Joseph, D.D., (2003). Power law
and composite power law friction factor correlations for laminar and turbulent gas-liquid
flow in horizontal pipelines. Int. J. Multiphase Flow 29, pp. 1605-1624.
Gaspari, G.P., Lombardi, C. and Peterlongo, G., (1964). Pressure drops in steam-water mixtures.
CISE Report R-83, Segrate, Italy.
Gill, L.E., Hewitt, G.F. and Lacey, P.M.C., (1964). Sampling probe studies of the gas core in
annular two-phase flow-II: Studies of the effect of phase flow rates on phase and velocity
distribution. Chem. Eng. Sci. 19, pp. 665-682.
Gill, L.E., Hewitt, G.F. and Lacey, P.M.C., (1965). Data on the upwards annular flow of air-water
mixtures. Chem. Eng. Sci. 20, pp. 71-88.
Hewitt, G.F. and Hall-Taylor, N.S., (1970). Annular Two-Phase Flow. Pergamon Press.
Hewitt, G.F. and Owen, D.G., (1987). Pressure Drop and Entrained Fraction in Fully Developed
Flow. Dataset No. 1. In Multiphase Science and Technology, Vol. 3, eds. Hewitt, G.F.,
Delhaye, J.M. and Zuber, N., Hemisphere, USA.
Two-Phase Pressure Drop 35

Hinkle, W.D., (1967). A study of liquid mass transport in annular air-water flow. Ph.D. Thesis,
Massachusetts Institute of Technology, Boston, USA.
Kanno, H., Han, Y., Saito, Y. and Shikazono, N., (2010). Measurement of liquid film thickness in
microtube annular flow. 14th International Heat Transfer Conference, paper No. IHTC14-
23176, August 8-13, Washington D.C., USA.
Lockhart, R.W. and Martinelli, R.C., (1949). Proposed correlation of data for isothermal two-phase,
two-component flow in pipes. Chem. Eng. Progr. 45, pp. 39-48.
Lombardi, C. and Carsana, C.G., (1992). A dimensionless pressure drop correlation for two-phase
mixtures flowing upflow in vertical ducts covering wide parameter range. Heat Technol. 10,
pp. 125-141.
Martinelli, R.C. and Nelson, D.B., (1948). Prediction of pressure drop during forced-circulation
boiling of water. Trans. ASME 70, pp. 695-702.
McAdams, W.H., Woods, W.K. and Heroman, L.C., (1942). Vaporization inside horizontal tubes
II-benzene-oil mixtures. Trans. ASME 64, pp. 193-200.
McAdams, W.H., (1954). Heat Transmission. McGraw Hill, New York, USA.
Mishima, K. and Hibiki, T., (1996). Some characteristics of air-water two-phase flow in small
diameter vertical tubes. Int. J. Multiphase Flow 22, pp. 703-712.
Müller-Steinhagen, H. and Heck, K., (1986). A simple friction pressure drop correlation for two-
phase flow in pipes. Chem. Eng. Process. 20, pp. 297-308.
Oliemans, R.V.A., Pots, B.F.M. and Trompé, N., (1986). Modeling of annular dispersed two-phase
flow in vertical pipes. Int. J. Multiphase Flow 12, pp. 711-732.
Ong, C.L., (2010). Macro-to-microchannel transition in two-phase flow and evaporation. Ph.D.
thesis, Swiss Federal Institute of Technology-EPFL, Lausanne, Switzerland.
Owens, W.L., (1961). Two-phase pressure gradient. ASME Int. Develop. Heat Transf. Part II, pp.
363-368.
Revellin, R. and Thome, J.R., (2007). Adiabatic two-phase frictional pressure drops in
microchannels. Exp. Therm. Fluid Sci. 31, pp. 673-685.
Sekoguchi, K., Nishikawa, K., Sato, Y. and Kariyasaki, A., (1968). Two-phase flow characteristics
in the disturbed flow region after mixing air and water. JSME Bull. 11, pp. 647-653.
Selander, W.N., (1978). Explicit formulas for the computation of friction factors in turbulent pipe
flow. Vol. 6354 of Atomic Energy of Canada Limited.
Shannak, B.A., (2008). Frictional pressure drop of gas liquid two-phase flow in pipes. Nuc. Eng.
Des. 238, pp. 3277-3284.
Shearer, C.J. and Nedderman, R.M., (1965). Pressure gradient and liquid film thickness in co-
current upwards flow of gas/liquid mixtures: application to film cooler design. Chem. Eng.
Sci. 20, pp. 671-683.
Shedd, T.A., (2010). Void fraction and pressure drop measurements for refrigerant R410a flows in
small diameter tubes. Preliminary AHRTI Report No. 20110-01.
Shiba, M. and Yamazaki, Y., (1967). A comparative study on the pressure drop of air-water flow.
JSME Bull. 10, pp. 290-298.
Silva Lima, R.J., Moreno Quiben, J. and Thome, J.R., (2009). Flow boiling in horizontal smooth
tubes: new heat transfer results for R134a at three saturation temperatures. Appl. Therm. Eng.
29, pp. 1289-1298.
Silvestri, M., Casagrande, I., Cravarolo, L., Hassid, A., Bertoletti, S., Lombardi, C., Peterlongo, G.,
Soldaini, G., Vella, G., Perona, G. and Sesini, R., (1963). A Research Program in Two-Phase
Flow. CISE Report, Segrate, Italy.
36 J.R. Thome & A. Cioncolini

Steiner, D., (1987). Pressure Drop in Horizontal Flows. Dataset No. 6. In Multiphase Science and
Technology, Vol. 3, eds. Hewitt, G.F., Delhaye, J.M. and Zuber, N., Hemisphere, USA.
Stephan, K. (1992). Heat Transfer in Condensation and Boiling. Springer-Werlag, Berlin,
Germany.
Sun, L. and Mishima, K., (2009). Evaluation analysis of prediction methods for two-phase flow
pressure drop in mini-channels. Int. J. Multiphase Flow 35, pp. 47-54.
Tibiriça, C.B. and Ribatski, G., (2011). Two-phase frictional pressure drop and flow boiling heat
transfer for R245fa in a 2.32 mm tube. Heat Transfer Eng. 32, pp. 1139-1149.
Woldesemayat, M.A. and Ghajar, A.J., (2007). Comparison of void fraction correlations for
different flow patterns in horizontal and upward inclined pipes. Int. J. Multiphase Flow 33,
pp. 347-370.
Würtz, J., (1978). An experimental and theoretical investigation of annular steam-water flow in
tubes and annuli at 30 and 90 bar. Risø National Laboratory, Report No. 372, Denmark.
Zhang, M. and Webb, R.L., (2001). Correlation of two-phase friction for refrigerants in small
diameter tubes. Exp. Therm. Fluid Sci. 25, pp. 131-139.

View publication stats

You might also like