0% found this document useful (0 votes)
40 views

Two-Phase Flow Pattern Maps For Macrochannels: October 2015

This chapter discusses two-phase flow patterns and flow pattern maps for macrochannels. It describes the common flow patterns observed in vertical tubes, including bubbly flow, slug flow, churn flow, annular flow, and mist flow. It also presents several widely used flow pattern maps for vertical tubes but notes that these maps do not account for thermal influences on flow pattern transitions during evaporation or condensation. Additionally, it describes flow patterns and maps for horizontal tubes where some maps include thermal effects on flow pattern changes.

Uploaded by

akinhaci
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
40 views

Two-Phase Flow Pattern Maps For Macrochannels: October 2015

This chapter discusses two-phase flow patterns and flow pattern maps for macrochannels. It describes the common flow patterns observed in vertical tubes, including bubbly flow, slug flow, churn flow, annular flow, and mist flow. It also presents several widely used flow pattern maps for vertical tubes but notes that these maps do not account for thermal influences on flow pattern transitions during evaporation or condensation. Additionally, it describes flow patterns and maps for horizontal tubes where some maps include thermal effects on flow pattern changes.

Uploaded by

akinhaci
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 48

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

net/publication/300569658

Two-Phase Flow Pattern Maps for Macrochannels

Chapter · October 2015


DOI: 10.1142/9789814623216_0019

CITATIONS READS

0 3,915

2 authors, including:

Andrea Cioncolini
The University of Manchester
85 PUBLICATIONS   1,057 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

Fluid dynamics of hairy coated profiles: experimental and numerical study 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 2

Two-Phase Flow Pattern Maps for Macrochannels

In this chapter, two-phase flow patterns and flow pattern maps in macrochannels
are presented in such detail so as to be useful for the other topics addressed in
this book. First, the flow patterns occurring inside vertical tubes are described
and several widely quoted flow pattern maps are presented, albeit none which
take into account any diabatic/evaporating/condensing influences on the flow
and its transition from one regime to another. Next, flow patterns and flow
pattern maps for two-phase flow in horizontal tubes are described, where in this
case several thermal influences of evaporation on the flow pattern transitions are
included in the making of diabatic flow pattern maps.

1. Introduction and Practical Relevance

The local distribution of the liquid and vapor flowing in a channel has an
important impact on heat transfer and pressure drop in vertical and horizontal
channels and tubes. The “geometric structures” of two-phase flows are
categorized into various regimes that are used to characterize such flows, which
can then be used to build flow regime specific models for flow boiling heat
transfer coefficients, two-phase pressure drops, void fractions, entrainment,
dryout, critical heat flux and so forth. In fact, the best and most reliable
mechanistically based prediction methods attempt to capture the two-phase flow
structure of the particular flow pattern to account for its dominant flow
phenomena: bubbles, liquid slugs, liquid films, entrained droplets, etc. To
simulate flow boiling along channels in which the flow pattern goes from one
flow regime to another from inlet to outlet, accurate flow pattern maps are thus
imperative to reliably identify what type of flow pattern exists at the local flow
conditions along the channel.
Typically, a tube depicts macrochannel flow characteristics when it is larger
than about 3 mm in internal diameter and below about 30 to 50 mm diameter.
Channels less than 3 mm tend to display some special characteristics that
differentiate them from macrochannels, such as the absence of flow stratification
in horizontal flows and some new two-phase flow structures. Even though large
2 J.R. Thome & A. Cioncolini

(> 50 mm) diameter pipes tend to function similar to what are here called
“macrochannels”, the vast majority of flow pattern maps developed from
experimental flow observations are taken in the range from 3 to 50 mm and thus
may not work (or have never been validated) when extrapolated to larger sizes.
Hence, the present chapter is considered to be appropriate for tubes and channels
with internal diameters of about 3-50 mm.

2. Description of Flow Patterns in Vertical Tubes

Co-current upflow and downflow of gas (or vapor) and liquid in a vertical tube is
characterized by recognizable flow structures and these “geometries” are used to
classify the flows into a number of two-phase flow regimes or flow patterns.
While many different names can be found in the literature, the most commonly
accepted vertical channel regimes are referred to as: bubbly flow, slug flow, churn
flow, annular flow and mist flow. These are depicted schematically in Fig. 1. The
individual flow patterns are described as follows:

Figure 1. Two-phase flow patterns in vertical upflow


Two-Phase Flow Pattern Maps for Macrochannels 3

Bubbly flow. In this regime, a large number of small bubbles, notably smaller
than the channel diameter, are observed in the continuous liquid phase. The
bubbles may vary widely in size and shape but they are typically rather densely
packed and flowing side-by-side and one after another, creating a very complex
flow. These bubbles interact with one another: colliding, coalescing, breaking up,
etc. During evaporation, this regime is normally the first one encountered by the
flow in the range of vapor qualities greater than or equal to 0.0. In fact, bubbly
flow is often already occurring in subcooled boiling before the fluid reaches its
saturation conditions.

Slug flow. This regime is characterized by large bubbles that span nearly the
entire channel cross-section and are separated by distinct liquid slugs, the latter
giving this regime its name. These bubbles have a cylindrical shape with a
hemispherical nose and a blunt fluctuating tail; they are commonly referred to as
Taylor bubbles after the instability of that name. The bubbles are typically about
1.5 to 3.0 times longer in length than the tube diameter. The flow of these long
bubbles trap a thin liquid film between themselves and the tube wall, which often
flows downward due to the force of gravity, even when the net flow of fluid is
upward in vertical up flow. The liquid slugs in between these long bubbles often
include some small bubbles, remaining from the prior bubbly flow regime or
created by vapor breaking off from the tails of the long bubbles.

Churn flow. This regime is a very chaotic fluctuating type of flow that is
passed through during the transition from a stable slug flow to a stable annular
flow. Essentially, the tails of the elongated bubbles begin to break up and create
large voids in the liquid slugs, which in turn become unstable themselves, and
which then in turn destabilize the noses of the trailing bubbles. Additionally, the
liquid film trapped between the long bubble and the wall may also become
unstable when a relative parity of the gravity and shear forces acting in opposing
directions on the thin liquid film is reached in vertical upward flow. Churn flow
is typically to be avoided in two-phase piping, because of the pressure and flow
oscillations that it induces. Some flow pattern maps treat this regime as part of
the transition region between slug flow and annular flow (in some ways
analogous to that between a laminar and turbulent flow in single-phase flow) and
hence do not cite it as a flow pattern.

Annular flow. This regime is named after the continuous annular liquid film
that flows on the internal perimeter of the channel with a higher velocity gas (or
vapor) core in the center. The interface of the annular liquid film is disturbed by
4 J.R. Thome & A. Cioncolini

high frequency waves and ripples. In addition, liquid droplets are usually
entrained in the gas core and small bubbles may be trapped in the annular film by
roll waves. This flow regime is particularly stable and is the desired flow pattern
in piping and at the exit of evaporator tubes. When the flow depicts entrained
droplets traveling in periodic coherent structures as clouds or wisps of liquid in
the central vapor core, some flow pattern maps refer to this as wispy annular
flow, but in general this is treaded for all practical purposes as an annular flow.

Mist flow. A mist flow has the gas (or vapor) as the continuous phase and
numerous very small droplets entrained in it, whose presence can often be
difficult to observe without the help of illumination and/or magnification. Mist
flow is encountered at high vapor qualities at the point where the annular film is
thinned by the shear of the gas core on the interface until it becomes unstable,
such that all the liquid is entrained as droplets in the continuous gas phase, or
during flow boiling when the critical heat flux has been surpassed. Impinging
liquid droplets intermittently wet the tube wall locally but heat transfer is
significantly decreased with respect to an annular flow.

Flow patterns in vertical downflows look similar to those in vertical upflows,


although there are some exceptions. For instance, at low flow rates in the slug
flow regime the nose of the bubble may be flat (due to the buoyancy forces
acting opposite to the surface tension) and the tail of the bubble may be round
(for the same reason). In annular downflows, the flow may also appear to be
simply a falling film, since in this case the interfacial shear and gravity forces
work together in the same direction.
Other regime names than those above, such as bubbly-slug flow, are often
utilized to describe transition zones between the above major regimes. Since
these transition zones can be quite broad in terms of the range of the measured
variables, some flow pattern maps segregate them into separate flow pattern(s).
Such distinctions impede the comparison of leading flow pattern maps against
one another since they have differing regime definitions. Similarly, it is also
difficult to validate leading flow pattern maps versus independent observations
that use different regime designations, as will be seen in the next section.

3. Adiabatic Flow Pattern Maps for Vertical Tubes

A two-phase flow pattern map is used to predict and identify the two-phase flow
regime that will occur based on the local flow parameters: tube diameter, mass
flow rates of the two phases and their respective fluid physical properties. If one
Two-Phase Flow Pattern Maps for Macrochannels 5

wishes, this can be considered as analogous to the Moody single-phase flow


diagram for flow in tubes in which one calculates the tubular Reynolds number
and then determines whether one is in laminar, transition or turbulent flow (and
the flow’s friction factor). In two-phase flows, the transitions are much more
complex phenomena because the two-phase flow structure is a function of the
flow and each phase can also be either laminar or turbulent. Hence, numerous
attempts have been made to prepare “flow pattern maps” for distinguishing
which flow regime will occur over what range of conditions and specifically
where the transitions between regimes occur.

Figure 2. Flow regimes in a vertical evaporator tube from Collier and Thome (1994) (flow is
upflow).
6 J.R. Thome & A. Cioncolini

The development of a flow pattern map first begins with some representative
database of flow pattern observations. Most often, this means tests with air and
water, which are the simplest to implement, while some databases also include
air (or other gases or steam) with oil and/or glycols. Working fluids in two-phase
systems other than water/steam are usually taken only for refrigerants and only
limited observations are usually available for other fluids of interest, such as
hydrocarbons and cryogenic fluids. Regarding flow pattern maps, many early
maps were developed based on plotting the observations versus various two-
phase parameters on log-log plots until a set of axes were found to segregate the
regimes into distinct zones, and then lines were drawn in at the boundaries
between the zones…that is, no prediction method was proposed for the
transitions, just lines or curves fit into such graphs. Other maps instead have
proposed non-dimensional expressions to predict the transitions, bringing more
physical properties into the map than just those included (if any) in the definition
of the two axes. More sophisticated maps have proposed mechanistic models of
the transition phenomena, and thus come closer to becoming a general use flow
pattern map, although even these still tend to be biased towards the fluids used in
their development. For a more fundamental treatment of two-phase flow
transitions than will be presented below, refer to the work of Barnea and Taitel
(1986) and a review paper presented by Taitel (1990).
For vertical evaporating upflows, Fig. 2 depicts the diagram of Collier and
Thome (1994), that shows the typical order of the flow regimes that would be
encountered from inlet to outlet from single-phase subcooled flow all the way
through to superheated vapor when a uniform heat flux is imposed on the tube.
Although this process is diabatic, the usual assumption is that adiabatic flow
pattern observations and maps are adequate for describing evaporating flows,
notwithstanding the influences of subcooled boiling, acceleration of the flow,
thinning of the film due to its evaporation, etc. Presently, no comprehensive
diabatic flow pattern maps for vertical flows are available that include such
effects, and hence the adiabatic assumption will be assumed here.
As depicted, the flow begins in the bubbly flow regime at the inlet at the point
of onset of nucleate boiling, which may begin in the subcooled zone or in the
saturated zone after the liquid has become superheated, depending on the local
wall superheating required to nucleate the bubbles in the superheated thermal
boundary layer on the heated tube wall. At subcooled conditions, such bubbles
tend to condense as they migrate into the subcooled core. After bubbly flow, the
slug flow regime is entered and then the annular flow regime with its
characteristic annular film of liquid. While the churn flow regime is not shown, it
is the transition path in between these two regimes. The annular film eventually
Two-Phase Flow Pattern Maps for Macrochannels 7

dries out due to evaporation, or the liquid film is hydrodynamically entrained by


the interfacial vapor shear, or the critical heat flux is surpassed, taking the flow
into the mist flow regime. The entrained liquid droplets in mist flow often persist
into the superheated vapor region, past the point of vapor quality x = 1.0.
To predict the local flow pattern in a tube, a graphical representation is used.
This is a diagram that displays the transition boundaries between the various flow
patterns and is typically plotted on a graph (i.e. a map) with log-log axes using
parameters to represent the liquid and gas velocities and some relevant physical
properties. Below, the most widely quoted flow pattern maps are presented.

Fair (1960) proposed the first widely quoted flow pattern map for vertical
upflow, illustrated in Fig. 3. It segregates the flows into four flow patterns,
without a churn flow regime that would normally lie in between the slug and
annular flow regions. The vertical axis is the total mass flux of liquid plus vapor
while its x-axis includes a vapor quality term and ratios of the phase densities and
dynamic viscosities.

Figure 3. Two-phase flow pattern map of Fair (1960) for vertical tubes

This map was originally proposed in US units (G in lbft-2 s-1) but is given here
in SI units that are now more commonly used. To utilize this map, one first
calculates the value of the x-axis and the total mass flux G (kgm-2s-1) for the
8 J.R. Thome & A. Cioncolini

particular conditions and these values are then used to pinpoint if the flow is
bubbly, slug, annular or mist, where the dark lines show the transition thresholds
between the regimes. This map is thus dimensional on its vertical axis but
dimensionless on its horizontal axis. One should consider the transition curves on
flow pattern maps to be surrounded by a transition zone, analogous to that
between laminar and turbulent flows. For example, in the transitions between
bubbly (or bubble) flow and slug flow, one sees partially formed elongated
bubbles with a bubbly wake that are interspersed with bubbly flow regions, and
thus alternating with one another. To implement this map into a computer
simulation, the following expressions have been extracted from the graph of Fair
(1960) for his three transition lines (Y for vertical axis and X for horizontal axis):

For bubbly flow to slug flow:


Y = 80.6 X -1.492 where range is 0.2 ≤ X ≤ 3 and 15.6 ≤ Y ≤ 879 (1)

For slug flow to annular flow:


Y = 503 X -1.231 where range is 0.5 ≤ X ≤ 18 and 14.6 ≤ Y ≤ 1221 (2)

For annular flow to mist flow:


Y = 2988 X -1.044 where range is 3 ≤ X ≤ 90 and 28.3 ≤ Y ≤ 976 (3)

The two axis are as follows:


0.9 0.5 0.5
⎛ x ⎞ ⎛ ρl ⎞ ⎛ µg ⎞
X =⎜ ⎟ ⎜ ⎟ ⎜⎜ ⎟⎟ (4)
⎝1− x ⎠ ⎜ρ ⎟
⎝ g ⎠ ⎝ µl ⎠
Y = G (kgm−2 s −1 ) (5)

Hewitt and Roberts (1969) proposed another widely quoted map using the
square of mass flux of the gas flow as the vertical axis and the square of mass
flux of the liquid flow as the horizontal axis, divided by their respective densities,
as shown in Fig. 4. They divided their flow observations into bubbly, bubbly-
slug/slug, churn, annular and wispy annular regions. The mass fluxes of the
liquid Gl and gas (or vapor) Gg are calculated using the local vapor quality x, the
values of the x and y coordinates are determined, and the intersection of the two
values on the map identifies the flow pattern predicted to exist. The mass flux of
Two-Phase Flow Pattern Maps for Macrochannels 9

the liquid Gl and the mass flux of the gas (or vapor) Gg are defined respectively
as:
Gl = (1 − x)G; Gg = xG (6)

where G is the total mass flux of the liquid and vapor, i.e. the total mass flow rate
divided by the channel flow area. In general, other researchers do not divide
annular flow into two parts, and hence this map’s wispy annular flow regime is
usually treated as an annular flow. Furthermore, in this map there is no distinct
line to divide slug flow and bubbly-slug flow, and hence these are grouped
together into one regime, making it difficult to compare this map’s transition
boundaries versus other maps, such as that of Fair (1960), which itself does not
have a churn flow regime.

Figure 4. Two-phase flow pattern map of Hewitt and Roberts (1969) for vertical tubes.

To implement the Hewitt and Roberts (1969) map into a computer simulation,
the following expressions have been extracted from their graph for its transition
lines (Y is vertical and X is horizontal axis):

For bubbly flow to bubbly-slug flow:


Y = 5.36 104 X -1.075; where 2500 ≤ X ≤ 9000 and 1 ≤ Y ≤ 3 (7a)
11 -2.71
Y = 2 10 X ; where 600 ≤ X ≤ 900 and 3 ≤ Y ≤ 9 (7b)
10 J.R. Thome & A. Cioncolini

Y = 6000; where 9 ≤ X ≤ 90 (7c)

For slug flow to churn flow:


Y = 0.021 X 1.278; where 10 ≤ X ≤ 90 and 0.4 ≤ Y ≤ 6.8 (8a)
Y = 0.092 X 0.963; where 90 ≤ X ≤ 500 and 6.8 ≤ Y ≤ 35 (8b)
0.4489
Y = 2.18 X ; where 500 ≤ X ≤ 1000 and 35 ≤ Y ≤ 48 (8c)
0.051
Y = 33.4 X ; where 1000 ≤ X ≤ 6000 and 48 ≤ Y ≤ 52.5 (8d)

For churn flow to annular flow:


Y = 219 X -0.399; where 1.3 ≤ X ≤ 10 and 90 ≤ Y ≤ 200 (9a)
Y = 90; where 10 ≤ X ≤ 50000 (9b)

For annular flow to wispy annular flow:


X = 1000; where 90 ≤ Y ≤ 900 (10a)
-9 3.894
Y = 2 10 X ; where 1000 ≤ X ≤ 1300 and 900 ≤ Y ≤ 2500 (10b)
Y = 0.439 X 1.217; where 1300 ≤ X ≤ 3000 and 2500 ≤ Y ≤ 7000 (10c)

where:

Gl2
X= (kgm−1s − 2 ) (11)
ρl
Gg2
Y= (kgm−1s − 2 ) (12)
ρg

4. Description of Flow Patterns in Horizontal Tubes

The second important orientation for two-phase flows is horizontal channels. In


fact, two-phase flow patterns in horizontal tubes are similar to those in vertical
flows except for flow stratification effects. In a horizontal two-phase flow, the
liquid and vapor tend to separate due to buoyancy forces: liquid towards the
bottom of the tube and the gas towards the top. At sufficiently high flow rates the
buoyancy influence becomes negligible and the flows become very similar to
vertical flows. Flow patterns for co-current flow of gas and liquid in a horizontal
tube are shown in Fig. 5. The individual flow patterns are described as follows:
Two-Phase Flow Pattern Maps for Macrochannels 11

Bubbly flow. The numerous small gas bubbles are dispersed in the continuous
liquid-phase, however the bubbles tend to the upper half of the tube due to their
buoyancy. When shear forces are dominant at high liquid flow rates, the bubbles
tend to disperse uniformly in the tube. In horizontal flows, the bubbly flow
regime is typically only seen at high mass flow rates since at lower flow rates the
numerous bubbles in the upper half tend to quickly coalesce into a few large long
bubbles (that is, plug flow in the intermittent regime described below).

Figure 5. Two-phase flow patterns in horizontal flow (from Silva Lima, 2011).

Stratified flow. When the buoyancy forces are completely dominant at low
liquid and gas velocities, complete segregation of the two phases occurs: the gas
travels along the top of the channel and the liquid along the bottom. The interface
between the two phases is nearly flat without significant waves. Typically more
than half of the perimeter of the tube is dry while the remainder is wet.

Stratified-wavy flow. Increasing the gas velocity in a stratified flow,


interfacial shear on the liquid surface creates waves which not only travel in the
direction of flow but they also tend to wrap up around the perimeter towards the
12 J.R. Thome & A. Cioncolini

top of the tube. The amplitude of the waves is notable; however, their crests do
not reach the top of the tube. Typically a significant portion of the upper
perimeter of the tube remains dry, although an occasional large amplitude wave
may intermittently wet it.

Intermittent flow. Further increasing the gas velocity, the interfacial waves
become consistently large enough to wet the top perimeter, typically leaving
behind a liquid film such that the entire perimeter of the tube is nearly always
wet. The large amplitude waves are separated by smaller amplitude waves in
between that do not reach the top of the channel. Entrained bubbles are usually
seen in the large waves, having been trapped by the rolling motion of the waves.
The liquid slugs separating such large elongated bubbles could also be described
as large periodic amplitude waves and hence gives the name of this regime:
intermittent flow. Intermittent flow is sometimes subdivided into two other
regimes: plug flow and slug flow:

Plug flow. This flow regime is a liquid flow with large elongated gas bubbles
flowing along the top of the channel. The diameters of the elongated bubbles are
smaller than the tube such that the liquid phase is continuous.

Slug flow. When the gas flow rate is increased, the diameters of elongated
bubbles become similar to that of the channel and the flow begins to look similar
to slug flow in vertical channels. Typically the liquid film between the top of the
bubble and the channel is thin compared to that beneath the bubble.

Annular flow. Similar to vertical flows, when increasing the gas flow rate the
liquid forms a continuous annular film around the perimeter of the tube;
however, the liquid film is thicker at the bottom of the pipe than at the top until at
even higher gas flow rates the annular liquid ring becomes uniform in thickness.
The interface between the liquid annulus and the vapor core is disturbed by
waves and droplets may be dispersed in the gas core. At high gas fractions, the
top of the tube with its thinner film becomes dry first, so that the annular film
covers only the lower tube perimeter and thus this may also be classified as a
stratified-wavy flow.

Mist flow. Similar to vertical flow, at very high gas velocities, all the liquid
may be stripped from the wall and entrained as small droplets in the now
continuous gas phase. This tends to be a progressive process that begins first at
the top of the tube where the liquid film is thinner and then proceeds downstream
until all of the liquid is entrained as droplets. Sometimes rivulets of liquid are
Two-Phase Flow Pattern Maps for Macrochannels 13

also seen after the breakdown of the liquid film, i.e. liquid that has not yet been
entrained.

5. Adiabatic Flow Pattern Maps for Horizontal Tubes

Baker (1954) proposed one of the earliest and still widely quoted flow pattern
maps for predicting the transition between two-phase flow regimes for adiabatic
flow in horizontal tubes, depicted in Fig. 6.

Figure 6. Two-phase flow pattern map of Baker (1954) for horizontal tubes.

This map segregates the flows into bubbly (with a froth zone), plug, slug,
stratified, wavy (i.e. stratified-wavy), annular and dispersed (apparently either
annular flow with substantial entrainment or mist flow). To utilize the map, first
the liquid and gas mass fluxes are calculated. Then, the physical property ratio
parameters λ and ψ are determined as:
1/ 2 2 1/ 3
⎛ρ ρl ⎞ σ
⎟⎟ ; ψ = ⎛⎜ water ⎞ ⎡⎛⎜ µl ⎞⎛ ρ water ⎞ ⎤
λ = ⎜⎜ g ⎟ ⎢⎜ ⎟⎟⎜⎜ ⎟⎟ ⎥ (13)
⎝ ρ air ρ water ⎠ ⎝ σ ⎠ ⎢⎣⎝ µ water ⎠⎝ ρl ⎠ ⎥⎦
14 J.R. Thome & A. Cioncolini

The properties of the fluid at the desired operating condition are input for ρg,
ρl, µl and σ while the reference properties for water and air to be used are:

• ρwater = 1000 kgm-3;


• ρair = 1.23 kgm-3;
• µwater = 0.001 kgm-1s-1;
• σwater = 0.072 Nm-1.

The values of the x-axis and y-axis are then determined to identify the
particular flow regime. As can be seen, the physical property ratios are used to
apply the map at other operating conditions relative to the base case of air and
water at his standard conditions. To compare to other maps and designations,
usually the plug and slug regimes would be jointed together and called
intermittent flow while the froth flow would be ignored. To implement this map
into a computer simulation, the following expressions have been extracted from
their graph for its transition lines:

For stratified flow to wavy flow:

Y = -0.121 X + 9.403; where 5 ≤ X ≤ 26.3 and 6.2 ≤ Y ≤ 8.8 (14a)


Y = -0.092 X + 8.387; where 26.3 ≤ X ≤ 66.6 and 2.5 ≤ Y ≤ 6.6 (14b)

For wavy/stratified flow to plug/slug/annular flow:


Y = 1.52 104 X -2.082
; where 11 ≤ X ≤ 300 and 0.1 ≤ Y ≤ 100 (15)

For plug flow to slug flow:


-0.243
Y = 3.512 X ; where 98.8 ≤ X ≤ 2213 and 0.5 ≤ Y ≤ 1.1 (16)

For slug flow to annular flow:


-0.848
Y = 214.1 X ; where 30.3 ≤ X ≤ 54.2 and 7.4 ≤ Y ≤ 12.1 (17a)
Y = 21.55 X -0.277; where 54.2 ≤ X ≤ 130.7 and 5.8 ≤ Y ≤ 7.4 (17b)
Y = 0.008 X + 4.652; where 130.7 ≤ X ≤ 868.5 and 5.8 ≤ Y ≤ 11.3 (17c)
Y = 0.006 X + 6.605; where 732.8 ≤ X ≤ 2975 and 10.3 ≤ Y ≤ 22.7 (17d)

For annular flow to dispersed flow:


Two-Phase Flow Pattern Maps for Macrochannels 15

Y = 1.168 104 X -1.032


; where 108.6 ≤ X ≤ 208 and 50 ≤ Y ≤ 100 (18a)
-0.255
Y = 188.5 X ; where 208 ≤ X ≤ 634.4 and 37.4 ≤ Y ≤ 50 (18b)
Y = 0.002 X + 34.6; where 634.4 ≤ X ≤ 462 and 36 ≤ Y ≤ 43.5 (18c)

For plug/slug/annular/dispersed flow to bubbly/froth flow:


Y = 55.83 ln(X) - 427; where 2383 ≤ X ≤ 1.2 104 and 0.25 ≤ Y ≤ 100 (19a)
Y = 4 1043 X -13.08
; where 2133 ≤ X ≤ 2625 and 0.1 ≤ Y ≤ 1.9 (19b)

where:

X = Gl ψ (kgm−2 s −1 ) (20)

Y = Gg λ−1 (kgm−2 s −1 ) (21)

Taitel and Dukler (1976) proposed perhaps the mostly widely quoted map for
horizontal flow in tubes, depicted in Fig. 7 (left), based on their mechanistic
analysis of the flow transition mechanisms together with a selection of non-
dimensional parameters. The map uses the Martinelli parameter X, the gas
Froude number Frg and the parameters T and K. Their map is composed of three
graphs (frequently represented as a single graph, however). The Martinelli
parameter and the gas Froude number are:
1/ 2
⎡ (dp / dz )l ⎤ Gg
X =⎢ ⎥ ; Frg = (22)
⎢⎣ (dp / dz )g ⎥⎦ [
ρ g (ρl − ρ g )d g 1 / 2 ]
The parameters T and K are:
1/ 2
⎡ (dp / dz )l ⎤ 0.5
T =⎢ ⎥ ; K = Frg Rel (23)
⎢⎣ g (ρl − ρ g )⎥⎦
where g is the acceleration due to gravity, while the liquid-phase and vapor-
phase Reynolds numbers are:
Gl d Gg d
Rel = ; Re g = (24)
µl µg
The pressure gradient of the flow for phase k (where k is either l or g) is:
16 J.R. Thome & A. Cioncolini

2ƒ k Gk2
(dp / dz )k =− (25)
ρk d
For Rek < 2000, the laminar flow friction factor equation is used:
16
ƒk = (26)
Re k
For Rek > 2000, the turbulent flow friction factor equation is used (even for the
transition regime from 2000 to 10,000):

0.079
ƒk = (27)
Re1k/ 4

Taitel & Dukler: Updated


ANNULAR DISPERSED
SYMMETRIC BUBBLY

1
10
Froude Number

ANNULAR
ASYMMETRIC

0
10
STRATIFIED
WAVY INTERMITTENT

-1
10 -2 -1 0 1
10 10 10 10
Martinelli Parameter
Two-Phase Flow Pattern Maps for Macrochannels 17

Figure 7. Two-phase flow pattern map of Taitel and Dukler (1976) for horizontal tubes (top-left),
updated version according to Cioncolini and Thome (2013) (top-right) and measured asymmetric
liquid film profiles (bottom).

To implement the map, one first determines the Martinelli parameter X and
Frg. Using these two parameters on the top graph, if their coordinates fall in the
annular flow regime, then the flow pattern is annular. If the coordinates of Frg
and X fall in the lower left zone of the top graph, then K is calculated. Using K
and X in the middle graph, the flow regime is identified as either stratified-wavy
or as fully stratified. If the coordinates of Frg and X fall in the right zone on the
top graph, then T is calculated. Using T and X in the bottom graph, the flow
regime is identified as either bubbly flow or intermittent (plug or slug) flow. To
implement this map into a computer simulation, the following expressions have
been extracted from their graph for its transition lines:

For intermittent flow to bubbly flow:


-0.070
Y = 1.056 X ; where 2 ≤ X ≤ 50 and 0.8 ≤ Y ≤ 1 (28a)
Y = -0.116 ln(X) + 1.283; where 50 ≤ X ≤ 9000 and 0.2 ≤ Y ≤ 0.8 (28b)

For stratified flow to wavy flow:


Y = 1.405 ln(X) + 7.521; where 0.009 ≤ X ≤ 0.7 and 1 ≤ Y ≤ 6.9 (29a)
Y = -1.337 ln(X) + 6.620; where 0.7 ≤ X ≤ 70 and 1.0 ≤ Y ≤ 6.9 (29b)

For wavy flow to annular flow:


-0.077
Y = 0.813 X ; where 0.001 ≤ X ≤ 0.05 and 0.95 ≤ Y ≤ 1.3 (30a)
Y = -0.214 ln(X) + 0.305; where 0.05 ≤ X ≤ 2 and 0.17 ≤ Y ≤ 0.95 (30b)
-1.188
Y = 0.412 X ; where 2 ≤ X ≤ 12 and 0.02 ≤ Y ≤ 0.17 (30c)
-1.861
Y = 2.041 X ; where 12 ≤ X ≤ 60 and 0.001 ≤ Y ≤ 0.02 (30d)
18 J.R. Thome & A. Cioncolini

For bubbly/intermittent flow to annular flow:


X = 2; where 0.18 ≤ Y ≤ 0.10 (31)

where X is the Martinelli parameter, while Y stands for K, T or Frg depending on


the flow pattern as explained above. One should note that representations of this
graph sometimes include a “1/4” on the vertical axis of the T parameter but not
on the parameter T itself; the present equations ignore this and no fourth root of
T is used in its presentation or in the calculations here.
The flow map of Taitel and Dukler (1976) has been recently extended by
Cioncolini and Thome (2013) to explicitly account for symmetric and
asymmetric annular flow. In the lower range of horizontal annular flow, when the
flow conditions are appropriate, the liquid in the film tends to drain down the
tube walls under the action of gravity, giving rise to an asymmetric liquid film
distribution in the channel cross section characterized by a thin liquid film at the
channel top and a thicker liquid film at the channel bottom. In particular, the
liquid film at the channel bottom can be up to 100 times thicker than the liquid
film at the channel top, based on available data. The prediction of the degree of
circumferential asymmetry in the liquid film is critical for properly designing
evaporators and condensers with horizontal tubes. Their updated Taitel and
Dukler flow map is depicted in Fig. 7 (right). As can be seen, only the graph
specific to annular flow has been updated, while the other two graphs are
unchanged. In particular, the liquid film distribution in the channel cross section
is symmetric when the gas Froude number is above Frg ≈ 10. An example of
asymmetric liquid film thickness profiles is shown in Fig. 7 (bottom), which
refers to air-water flow in a 9.85 mm diameter tube (further details in Cioncolini
and Thome (2013)). According to the authors, the degree of asymmetry of the
liquid film profile can be predicted as follows:
1.45
0.366 Frg
ta tb−1 = 1.45
; Frg > 1 (32)
1 + 0.366 Frg
where ta is the average liquid film thickness while tb is the liquid film
thickness at the channel bottom.

The above flow pattern maps were all developed for adiabatic two-phase
flows but are often extrapolated for use with the diabatic processes of
evaporation or condensation.
Two-Phase Flow Pattern Maps for Macrochannels 19

6. Flow Pattern Maps for Evaporation in Horizontal Tubes

For evaporation in horizontal tubes, Fig. 8 from Collier and Thome (1994)
depicts the typical flow regimes, including cross-sectional views of the flow
structure. For condensation, the flow regimes are similar with the exception that
the top tube wall is not dry in stratified types of flow but instead is coated with a
thin condensing film of condensate.

Figure 8. Flow patterns during evaporation in a horizontal tube, from Collier and Thome (1994)
(flow is left to right).

Figure 9. Kattan-Thome-Favrat (1998a) flow pattern map at three boiling heat fluxes compared to
20 J.R. Thome & A. Cioncolini

adiabatic boundaries of Steiner (1993) for R-410A at 5 °C in a 13.84 mm internal diameter tube.

For small diameter tubes typical of heat exchangers, Kattan et al. (1998a,
1998b, 1998c) proposed a modification of the Steiner (1993) map, which itself is
a modification of the above Taitel and Dukler (1976) map. In particular, they
included a heat flux dependent method for predicting the onset of dryout at the
top of the tube in evaporating annular flows to make this into a diabatic and
adiabatic flow pattern map. This flow pattern map will be presented here as it is
used in Chapter 7 for predicting local flow boiling coefficients based on the local
flow pattern. The flow regime transition boundaries of the Kattan et al. flow
pattern map are depicted in Fig. 9 (bubbly flow is at very high mass velocities
and is not shown), comparing them to the previous boundaries of Steiner (1993).
The Kattan et al. map deviates from other maps in that it provides the transition
boundaries on a linear-linear graph with mass velocity plotted versus gas or
vapor quality for the particular fluid and flow channel, which is much easier to
use than the log-log format of other maps. Note that the rise in the mist flow
transition line at high vapor qualities in the Steiner map, observable when
plotting it on a G vs. x basis, predicts that a flow could pass from annular flow
into mist flow and then back into annular flow (following a horizontal line on the
plot), which is not physically realistic.
The transition equations for calculating the transition mass velocities are
given below. First of all, the transition boundary curve for Gwavy between annular
and intermittent flows to stratified-wavy flow is:
0.5
⎧ 16 A3gd g d ρ l ρ g ⎫
⎪ 0.5 ⎪ (33)
G =⎨
2
(
⎪ x 2π 1−(2hld −1) 2 ) ⎪
+ 50
wavy
− F2 (q) ⎤

⎪⎡ π 2 − F1 (q)⎛ We ⎞ ⎪
⎪⎢ 25 2 (1− x) ⎜ ⎟ +1⎥⎪
⎩⎢⎣ hld ⎝ Fr ⎠l ⎥⎦⎭
The high vapor quality portion of this curve depends on the ratio of the
Froude number (Frl) to the Weber number (Wel), where Frl is the ratio of the
inertia to the gravity forces while Wel is the ratio of inertia to surface tension
forces. The mass velocity threshold for the transition for Gmist from annular flow
to mist flow is:
0.5
⎧⎪ 7680A2gd g d ρ l ρ g ⎛ Fr ⎞ ⎫⎪ (34)
G =⎨
mist
2 2ξ
⎜ ⎟⎬
⎪⎩ x π Ph ⎝ We ⎠l ⎪⎭
Two-Phase Flow Pattern Maps for Macrochannels 21

Evaluating the above expression for the minimum mass velocity of the mist
flow transition gives the value of xmin, after which for the transition line becomes
flat for x > xmin:
Gmist = Gmin (35)

The transition between stratified-wavy flow and fully stratified flow is given
by the expression for Gstrat:
1/ 3

(
⎧⎪ (226.3)2 Ald A2gd ρ g ρ l − ρ g µ l g ⎫⎪ ) (36)
G =⎨strat
2 3 ⎬
⎪⎩ x (1− x)π ⎪⎭
The transition threshold into bubbly flow is given by the following expression
for Gbubbly:
1 / 1.75
2 1.25
⎧⎪ 256 Agd Ald (
d ρ l ρ l − ρ g g ⎫⎪ ) (37)
G bubbly
=⎨
1.75 0.25 ⎬
⎪⎩ 0.3164(1− x) π 2 Pid µ l ⎪⎭
In the above equations, the ratio of Wel to Frl is:

⎛ We ⎞ = g d ρ2
l (38)
⎜ ⎟
⎝ Fr ⎠ l
σ

and the friction factor is:


−2
⎡ ⎛ π ⎞⎤
ξ Ph = ⎢1.138 + 2log⎜⎜ 1.5 ⎟⎟⎥ (39)
⎣⎢ ⎝ Ald ⎠⎦⎥
The non-dimensional empirical exponents F1(q) and F2(q) in the Gwavy
boundary equation include the effect of heat flux on the onset of dryout of the
annular film determined from their flow boiling experiments and flow
observations, i.e. the transition of annular flow into annular flow with partial
dryout, the latter which is classified as stratified-wavy flow by the map. They are
calculated as follows:
2
⎛ q ⎞ ⎛ q ⎞
F (q) = 646.0⎜⎜ ⎟ + 64.8⎜ ⎟
1
⎟ ⎜q ⎟ (40)
⎝ q DNB ⎠ ⎝ DNB ⎠
⎛ q ⎞
F (q ) = 18.8⎜ ⎟+1.023
2 ⎜q ⎟
⎝ DNB ⎠
22 J.R. Thome & A. Cioncolini

The Kutateladze (1948) correlation for the heat flux of departure from
nucleate boiling, qDNB is used to normalize the local heat flux:

q DNB
1/ 2
= 0.131ρ g i gl [g(ρ − ρ )σ ]
l g
1/ 4
(41)

The vertical boundary between intermittent flow and annular flow is assumed
to occur at a fixed value of the Martinelli parameter Xtt equal to 0.34, where Xtt is
defined as:
0.5 0.125
0.875
⎛ 1− x ⎞ ⎛⎜ ρ g ⎞⎟ ⎛µ ⎞ (42)
X ⎜
= ⎟ ⎜ l⎟
⎝ x ⎠ ⎜⎝ ρ l ⎟⎠
tt
⎜ µg ⎟
⎝ ⎠
Solving for x, the vertical threshold line of the intermittent-to-annular flow
transition at xIA is:
−1
⎧⎡ −1 / 1.75 −1 / 7
⎪ ⎛ ρg ⎞ ⎛ µ ⎞ ⎤ ⎫⎪ (43)
xIA = ⎨⎢0.2914⎜ ⎟ ⎜ l ⎟ ⎥+1⎬
⎜ ⎟ ⎜ µg ⎟ ⎥
⎪⎩⎢⎣ ⎝ ρl ⎠ ⎝ ⎠ ⎦ ⎪⎭

Figure 10. Geometrical description of a stratified flow in a circular tube.

Figure 10 defines the geometrical dimensions of the flow where Pl is the


wetted perimeter of the tube, Pg is the dry perimeter in contact with only vapor, h
is the height of the completely stratified liquid layer, and Pi is the length of the
Two-Phase Flow Pattern Maps for Macrochannels 23

phase interface. Similarly Al and Ag are the corresponding cross-sectional areas of


the liquid and vapor. Normalizing with the tube internal diameter d, six
dimensionless variables are obtained:
h
P P, P P, g
h ld
= ,
d ld
=
d
l
gd
=
d (44)
P, A A, A A g
P id
=
d
i
ld
=
d2
l
gd
=
d2
For hld ≤ 0.5:
0.5 0 .5
P ld
= ⎛⎜ 8 hld

( ) −2 (hld (1− hld )) ⎞⎟ 3

P gd
= π − Pld (45)
0 .5 0.5
A ld
= ⎛⎜12

(hld (1− hld )) +8 (hld ) ⎞⎟
⎠hld
15

π
A gd
=
4
− A ld

For hld > 0.5:


0.5 0.5
P gd

(
= ⎛⎜ 8 1− hld ) −2 (hld (1− hld )) ⎞⎟ 3

(46)
P ld
= π − P gd
0. 5 0 .5
A gd
= ⎛⎜12

(hld (1− hld )) + 8(1− hld ) ⎞⎟⎠(1 − h ) 15 ld

π
A ld
=
4
− Agd

For 0 ≤ hld ≤ 1:

(hld (1−hld )) 0.5


(47)
P id
=2

Since h is unknown, an iterative method utilizing the following equation is


necessary to calculate the reference liquid level hld:
14
⎡ ⎤
⎢⎛ P gd + Pid ⎞ ⎛⎜ π ⎞⎟⎥
2

⎢⎜ ⎟ ⎜ 64 2 ⎟⎥ (48)
14

⎠ ⎝ Agd ⎠⎥⎥ ⎛⎜ π ⎞ 3

⎟⎟ ⎜⎜ A
π ⎛ 64 ⎞
2 ⎢⎝
X tt ⎢
=
⎜ 2
ld ⎟
Pld ⎠ ⎝ π P

⎢ ⎛ P gd + Pid Pid ⎞ ⎥⎝ ld ⎠

⎢⎜ ⎜ + ⎟ ⎥
⎣⎢ ⎝ Agd Ald ⎟⎠ ⎦⎥

Once the reference liquid level hld is known, the dimensionless variables
above are calculated and the transition curves for the new flow pattern map are
then determined.
24 J.R. Thome & A. Cioncolini

This map was developed from a database for five refrigerants: two single-
component fluids (R-134a and R-123), two near-azeotropic mixtures (R-402A
and R-404A) and one azeotropic mixture (R-502). The test conditions covered
the following range of variables: mass flow rates from 100 to 500 kgm-2s-1, vapor
qualities from 4-100%, heat fluxes from 440 to 36500 Wm-2, saturation pressures
from 0.112 to 0.888 MPa, Weber numbers from 1.1 to 234.5, and liquid Froude
numbers from 0.037 to 1.36. The Kattan et al. (1998a, 1998b, 1998c) flow
pattern map correctly identified 96.2% of these flow pattern data.
Zürcher et al. (1997) obtained additional two-phase flow pattern observations
for the zeotropic refrigerant mixture R-407C at an inlet saturation pressure of
0.645 MPa and the map accurately identified these new flow pattern data.
Zürcher et al. (1999) then also obtained two-phase flow pattern data for ammonia
flowing a 14 mm bore sight glass for mass velocities from 20 to 140 kgm-2s-1,
vapor qualities from 1-99% and heat fluxes in the upstream evaporator tube of
the same diameter from 5000 to 58000 Wm-2, all taken at a saturation
temperature of 4 °C and saturation pressure of 0.497 MPa. Thus, the mass
velocity range in the database was extended from 100 kgm-2s-1 down to 20 kgm-
2 -1
s . In particular, it was observed that the transition curve Gstrat was too low and
its expression was empirically corrected by adding +20x as follows:
13
⎧⎪ (226.3)2 Ald Agd
2
ρ g (ρl − ρ g )µl g ⎫⎪
Gstrat =⎨ ⎬ + 20 x (49)
⎪⎩ x (1 − x )π 3
2
⎪⎭
where Gstrat is in kgm-2s-1. The transition from stratified-wavy flow to annular
flow at high vapor qualities was instead observed to be too high and hence an
additional empirical term with an exponential factor modifying the boundary at
high vapor qualities was added to take this into account as:
2
(
⎛ x 2 −0.97
−⎜
) ⎞⎟
⎜ x (1− x ) ⎟
Gwavy ( new ) = Gwavy − 75e ⎝ ⎠
(50)

where the mass velocity is in kgm-2s-1. The movement of these boundaries has a
direct effect on the dry angle calculation θdry in the Kattan et al. (1998c) flow
boiling heat transfer model and shifts the onset of dryout to slightly higher vapor
qualities, which is in agreement with the ammonia heat transfer test data.
In addition, Zürcher et al. (1999) found that the onset of dryout effect in the
Kattan et al. (1998a, 1998b, 1998c) map was too strong compared to their new
extensive observations for ammonia. They recommended reducing the influence
by one-half, so the value of q in their expressions should be replaced with q/2.
Two-Phase Flow Pattern Maps for Macrochannels 25

To utilize this map, the following parameters are required: vapor quality (x),
mass velocity (G), tube internal diameter (d), heat flux (q), liquid density (ρl),
vapor density (ρg), liquid dynamic viscosity (µl), vapor dynamic viscosity (µg),
surface tension (σ), and latent heat of vaporization (igl), all in SI units. The local
flow pattern is identified by the following procedure:

1. Solve Eq. (48) iteratively with Eqs. (42),(45),(46) and (47);


2. Evaluate Eq. (44);
3. Evaluate Eqs. (38),(39),(40) and (41);
4. Evaluate Eqs. (33), (34) or (35), (36), (37) and (43);
5. Compare these values to the given values of x and m! to identify the flow
pattern.

Note that Eq. (50) should be used in place of Eq. (33) and Eq. (49) should be
used in place of Eq. (43) to utilize the most updated version. The map is thus
specific to the fluid properties, flow conditions (heat flux) and tube internal
diameter input into the equations. The map can be programmed into any
computer language, evaluating the transition curves in incremental steps of 0.01
in vapor quality to obtain a tabular set of threshold boundary points, which can
then displayed as a complete map with G vs. x as coordinates.

Zürcher-Favrat-Thome map. Following the above work, based on extensive


flow pattern observations for ammonia at 5 °C in a 14.0 mm horizontal sight
glass tube at the exit of the same diameter of evaporator tube, Zürcher et al.
(2002) proposed a new version of the transition boundary curve between annular
and intermittent flows to stratified-wavy flow, i.e. for Eq. (33), based on an
analysis of the dissipation effects in a two-phase flow. The interested reader is
referred to their paper for the details.

Thome-El Hajal map. For practicality for programming the flow pattern
map into heat exchanger design codes, and for consistency between the flow map
and the heat transfer model, an easier to implement version of the Kattan-Thome-
Favrat map was proposed by Thome and El Hajal (2003). In the previously
presented flow pattern map, the dimensionless variables Ald, Agd, hld and Pid were
calculated in an iterative procedure using the stratified flow void fraction model
illustrated in Fig. 10 and coming from the original Taitel and Dukler (1976) map.
On the other hand, the flow boiling heat transfer model of Kattan et al. (1998c)
uses the Steiner (1993) version of the Rouhani and Axelsson (1970) drift flux
model for horizontal tubes for the cross-sectional void fraction ε (see Chapter 4):
26 J.R. Thome & A. Cioncolini

−1

ε=
x ⎡
⎢(1 + 0.12 (1 − x ))⎜ + ⎟+
[
⎛ x 1 − x ⎞ 1.18 (1 − x ) gσ (ρl − ρ g ) 0.25 ⎤

] (51)
ρ g ⎢⎣ ⎜ρ
⎝ g ρ l

⎠ Gρl0.5 ⎥⎦
This drift flux model is easy to apply and gives the void fraction as an explicit
function of total mass velocity, whereas the iterative method of Taitel and Dukler
(1976) used previously does not. Hence, to be consistent, it makes sense to use
the same void fraction model in both the flow pattern map and the flow boiling
heat transfer model. For this, the Rouhani-Axelson model is a good choice as the
general method, at least for refrigerants, which has been proven experimentally
by 238 void fraction measurements for R-22 and R-410A made by Ursenbacher
et al. (2004) for stratified-wavy and intermittent types of flow. Using this void
fraction model, the values Ald and Agd are now directly determinable by first
calculating the void fraction and then going to their definitions:
A(1 − ε ) Aε
Ald = ; Agd = (52)
d2 d2
Geometrically, the dimensionless liquid height hld and the dimensionless
length of the liquid interface Pid can be expressed as a function of the stratified
angle θstrat (stratified angle around upper perimeter of the tube to stratified liquid
level):
⎡ ⎛ 2π − θ strat ⎞⎤ (53)
hld = 0.5⎢1 − cos⎜ ⎟⎥
⎣ ⎝ 2 ⎠⎦
⎛ 2π − θ strat ⎞ (54)
Pid = sin ⎜ ⎟
⎝ 2 ⎠
To avoid an iteration to calculate the value of θstrat, the geometrical expression
for the stratified angle θstrat is calculated from an approximate expression
proposed by Biberg (1999), evaluated in terms of void fraction, as follows:
1/ 3
⎧ ⎛ 3π ⎞ ⎫
[ 1/ 3
⎪π (1 − ε ) + ⎜ ⎟ 1 − 2(1 − ε ) + (1 − ε ) − ε ⎪
⎪ ⎝ 2 ⎠
1/ 3

]
θ strat = 2π − 2⎨ (55)

⎪− 1 (1 − ε )ε [1 − 2(1 − ε )]1 + 4 (1 − ε )2 + ε 2 ⎪
[ ( )]
⎪⎩ 200 ⎪⎭

As the void fraction is a function of mass velocity, it influences the position


of the transition curves that involve ε in the Thome-El Hajal map. The strongest
effects of mass velocity are typically observed on the SW-I/A transition curve for
vapor qualities below 0.1 and at very low mass velocities, where the transition
Two-Phase Flow Pattern Maps for Macrochannels 27

curve goes up with increasing mass velocity, becoming less significant as the
vapor quality increases and at higher mass velocities. The boundary curve A-M
also moves up marginally with increasing mass velocity. Figure 11 shows this
new map of Thome and El Hajal (2003) compared to that of Zürcher et al.
(2002). In implementing the method for design purposes, the actual mass velocity
is used to calculate the transition curves while for expediency in calculating the
flow pattern maps below, a fixed value of mass velocity was used for evaluating
the entire map.

Figure 11. Flow pattern map of Thome and El Hajal (2003) compared to that of Zürcher et al.
(2002) for R-410A at 5°C in a 13.84 mm internal diameter tube with its equations evaluated at G =
200 kgm-2s-1 for several heat fluxes.

At this point, it is instructive to see some actual flow patterns. For this
purpose, some high quality photographs of two-phase flow patterns in horizontal
tubes are available from Barbieri et al. (2005) taken in the test facility described
in Barbieri and Sáiz-Jabardo (2006). They are for 500 kgm-2s-1 for R-134a in a
sight glass tube at the exit of an evaporator tube (note some external
condensation on the outside of the tube in some of the photographs). Figure 12
shows some photographs of stratified types of flow. Figure 13 shows the
sequence of events in the intermittent flow regime, characterized by a cyclic
variation between low amplitude waves (top photograph) and large amplitude
28 J.R. Thome & A. Cioncolini

waves (bottom photograph). Finally, Fig. 14 presents some images of annular


flow.
Two-Phase Flow Pattern Maps for Macrochannels 29

Figure 12. Stratified flow regime photographs of Barbieri et al. (2005). Top: stratified flow; middle
and bottom: stratified-wavy flow. Tube diameter: 15.8 mm.
30 J.R. Thome & A. Cioncolini

Figure 13. Intermittent flow regime photographs of Barbieri et al. (2005) with a sequence of
interfacial waves in a 9.52 mm tube (top) followed by climbing waves in a 15.8 mm tube (middle)
and then a large amplitude wave reaching the top of the channel in a 15.8 mm tube (bottom).
Two-Phase Flow Pattern Maps for Macrochannels 31

Figure 14. Annular flow regime photographs of Barbieri et al. (2005) showing three images of the
interfacial waves on an annular liquid film. Top: 9.52 mm tube; middle: 15.8 mm tube; bottom:
15.8 mm tube.
32 J.R. Thome & A. Cioncolini

Wojtan-Ursenbacher-Thome map. The flow pattern map of Kattan et al.


(1998a) was developed primarily for tests at vapor qualities higher than 0.15 and
also without benefit of a dedicated experimental campaign on the effect of heat
flux on the initiation and completion of dryout at high vapor qualities. In light of
the dynamic void fraction measurements made by Wojtan et al. (2004) and their
video observations for mass velocities between 70-200 kgm-2s-1 for a 13.84 mm
horizontal sight glass tube, the following conclusions were reached:

1. Fully stratified flow was not detected at any of the mass velocities tested;

2. In the vapor quality range 0 < x < xIA, an alternating flow structure of liquid
slugs and stratified-wavy interfaces was observed (where xIA is the vertical
line separating intermittent and annular flows);

3. The transition from slug/stratified-wavy flows to fully stratified-wavy flows


without any slugs appeared approximately to occur at xIA;

4. Only slug flow was observed for the zone identified by the Thome-El Hajal
map to be in the stratified-wavy region for G > Gwavy(xIA).

Based on these observations, the stratified-wavy flow region of the Thome-El


Hajal version of the flow map was modified by Wojtan et al. (2005a) as follows:

1. A new transition line was added at Gstrat = Gstrat(xIA) at x < xIA (this creates a
new horizontal transition line to the left of xIA and modifies the boundary of
the stratified (S) regime);

2. The stratified-wavy region has been divided into three subzones:


• For G > Gwavy (xIA), this becomes the Slug zone.
• For Gstrat < G < Gwavy(xIA) and 0 < x < xIA, this becomes the Slug/Stratified-
Wavy zone.
• For 1 > x ≥ xIA, this remains as the Stratified-Wavy zone.

Figure 15 depicts the new flow pattern map calculated for R-22 with the
above modifications applied to the Thome-El Hajal version of Kattan-Thome-
Favrat flow pattern map to better describe the actual character of the flow. The
dash lines correspond to the new dryout and mist flow transition curves described
below.
As depicted in Fig. 16, dryout occurs at the top of a horizontal tube first at xdi
(cross section A-A), where the annular liquid film is thinner, and then dryout
Two-Phase Flow Pattern Maps for Macrochannels 33

proceeds around the perimeter of the tube along its length (cross section B-B)
until reaching the bottom (cross section C-C) where the liquid film disappears at
xde. Thus, dryout in a horizontal tube takes place over a range of vapor qualities,
beginning as an annular flow and ending when the fully developed mist flow
regime is reached. This flow regime between xdi and xde is called dryout.

Figure 15. New flow pattern map simulated for R-22 at 5 °C in a 13.84 mm channel using 100 kgm-
2 -1
s and 2.1 kWm-2 to calculate the transition curves.

Figure 16. Dryout in a horizontal tube. (a) Dryout zone during evaporation in a horizontal tube
beginning at xdi at top of tube and ending at xde at bottom of tube; (b) Cross sections: A-A onset of
dryout in annular flow; B-B dryout; C-C end of dryout and beginning of mist flow.
34 J.R. Thome & A. Cioncolini

Since it is difficult to determine the onset and completion of dryout only from
visual observations in a sight glass tube, a large number of experimental flow
boiling heat transfer points were measured by Wojtan et al. (2005a) for R-22 and
R-410A at mass velocities from 70 to 700 kgm-2s-1 and heat fluxes from 2.0 to
57.5 kWm-2. The tube internal diameters tested were 13.84 mm for R-22 and R-
410A and an 8.00 mm tube for R-410A. These flow boiling heat transfer data
were used to identify the locations of xdi and xde. As illustrated in Fig. 17, the
sharp change in the heat transfer coefficient with increasing vapor quality
indicates the inception of dryout whereas the end of this decrease of heat transfer
coefficient marks the end of dryout and the beginning of mist flow. The
observations in the sight glass confirmed that the onsets of dryout and mist flow
appeared at the same vapor quality as detected by the heat transfer
measurements.

Figure 17. Experimental heat transfer coefficients in the 13.84 mm test section for R-22 at 5 oC
with an initial heat flux of 57.5 kWm-2 and mass flux of 600 kgm-2s-1.

Analyzing experimental results and observations in the sight glass, it is


obvious that usually there is no step-wise jump in the transition from annular
flow to mist flow. The first attempt to model the annular-dryout transition during
evaporation in horizontal tubes was made by Lavin and Young (1965). They
proposed a new transition between the annular and dryout zones based on the
Two-Phase Flow Pattern Maps for Macrochannels 35

Weber number for R-22 and R-12. Lavin and Young (1965) observed the dryout
process, but with the apparatus used, they could not obtain the heat transfer
coefficient within the dryout regime nor study the conditions under which the
dryout regime ends and a stable mist flow was established.
Since dryout occurs over an interval of vapor quality along the flow channel,
Mori et al. (2000) defined the inception of dryout to be xdi and the completion of
dryout to be xde and then used three characteristic regimes that they named S1, S2
and S3 to predict their values. The best agreement of the Wojtan et al. (2005a)
values of xdi and xde identified from their heat transfer data were given by their
regime S2, whose corresponding transition expressions by Mori et al. (2000) are:

xdi = 0.58e
[0.52−0.000021We 0.96 −0.02
g Frg (ρg ρl )−0.08 ]
(56)

xde = 0.61e
[0.57−0.0000265We 0.94 −0.02
g Frg (ρg ρl )−0.08 ]
(57)

These expressions of Mori et al. (2000) were modified by Wojtan et al.


(2005a) to include the heat flux effect observed from their results for R-22 and R-
410A evaporating at 5 °C in 8.00 and 13.84 mm diameter test sections for heat
fluxes up to 57.5 kWm-2, using the dimensionless heat flux ratio (q/qDNB) and new
empirical factors. Thus, the new limits for the beginning and end of the dryout
regime in their updated flow map are calculated with their following transition
equations:

xdi = 0.58e
[0.52−0.235We 0.17
g Frg
0.37
(ρg ρl )0.25 (q qDNB )0.7 ]
(58)

xde = 0.61e
[0.57−0.0058We0.38 0.15
g Frg (ρg ρl )−0.09 (q qDNB )0.27 ]
(59)

where qDNB is calculated with the expression in Eq. (41) of Kutateladze (1948).
After inversion of these two equations to solve for the mass velocity in terms of
vapor quality, the annular-to-dryout boundary (A-D) and the dryout-to-mist flow
boundary (D-M) transition equations for xdi and xde become respectively:
−0.17 0.926
⎡ 1 ⎛ 0.58 ⎞⎛ d ⎞ ⎤
⎛ ⎞
⎢ ⎜⎜ ln⎜ ⎟ + 0.52 ⎟⎟⎜
⎜ ⎟ ⎥
⎢ 0.235 ⎝ ⎝ x ⎠ ⎠⎝ ρ gσ ⎟⎠ ⎥ (60)
Gdryout =⎢ −0.37

−0.25 −0.7
⎢⎛ 1 ⎞ ⎛ ρg ⎞ ⎛ q ⎞ ⎥
⎢⎜⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎟ ⎥
⎢⎣⎝ gdρ g (ρ l − ρ g ) ⎟⎠ ⎝ ρ l ⎠ ⎝ qDNB ⎠ ⎥⎦
⎜ ⎟ ⎜
36 J.R. Thome & A. Cioncolini

−0.38 0.943
⎡ 1 ⎛ 0.61 ⎞⎛ d ⎞ ⎤
⎛ ⎞
⎢ ⎜⎜ ln⎜ ⎟ + 0 .57 ⎟⎜
⎟⎜ ρ σ ⎟
⎟ ⎥
⎢ 0.0058 ⎝ ⎝ x ⎠ ⎠⎝ g ⎠ ⎥ (61)
Gmist =⎢ −0.15

0.09 −0.27
⎢⎛ 1 ⎞ ⎛ ρg ⎞ ⎛ q ⎞ ⎥
⎢⎜⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎟ ⎥
⎣⎢⎝ gdρ g (ρ l − ρ g ) ⎠ ⎝ ρ l ⎠ ⎝ qDNB ⎠
⎟ ⎜ ⎟ ⎜
⎦⎥
Including the stratified-wavy region modifications together with the new A-D
and D-M transition curves into their map, the implementation procedure for the
Wojtan-Ursenbacher-Thome map is as follows:

1. The geometric parameters ε, Ald, Agd, hld, Pid and θstrat are calculated using
Eqs. (51)-(55);

2. As the effect of heat flux at high vapor quality is captured by the A-D
and D-M transition curves, the SW-I/A transition is first calculated from
the following adiabatic version of Eq. (33):
0.5
−1
⎧⎪ 16 A3 gdρ ρ ⎡ π2 ⎛ Wel ⎞ ⎤ ⎫⎪
Gwavy =⎨ gd l g
⎢ ⎜⎜ ⎟⎟ + 1⎥ ⎬ + 50 (62)
2 2
[
⎪⎩ x π 1 − (2hld − 1)
2
]0.5 2
⎢⎣ 25hld ⎝ Frl ⎠ ⎥⎦ ⎭⎪

3. The stratified-wavy region is then subdivided into three zones as follows:


i. G > Gwavy(xIA) gives the slug flow zone;
ii. Gstrat < G < Gwavy(xIA) and 0 < x < xIA give the Slug/Stratified-Wavy
zone;
iii. 1 > x ≥ xIA gives the Stratified-Wavy zone.

4. The S-SW transition is calculated from the original boundary expression


Eq. (36) but now Gstrat = Gstrat(xIA) when x < xIA, the latter which gives the
flat horizontal part of the boundary for 0 ≤ x ≤ xIA.

5. The I-A transition is calculated from the original boundary given by Eq.
(43) and extended down to its intersection with Gstrat.

6. The A-D boundary is calculated from Eq. (60) where its value takes
precedent over the value from step 2 above when its value is smaller than
Gwavy.
Two-Phase Flow Pattern Maps for Macrochannels 37

7. The D-M boundary is calculated from Eq. (61) but since the A-D and D-
M lines are not parallel these boundaries can intersect, so that when xde <
xdi then xde is set equal to the value of xdi and no D region exists (at high
mass velocities and low heat fluxes where this occurs, the high vapor
shear will tend to make the annular film to be of uniform thickness and
hence it seems reasonable that the entire perimeter becomes dry
simultaneously at xdi).

8. The following logic is applied to define the transitions in the high vapor
quality range for the onset of dryout in the map, referred to as Gdryout,
implemented in the following order:
• If Gstrat(x) ≥ Gdryout(x), then Gdryout = Gstrat(x);
• If Gwavy(x) ≥ Gdryout(x), then Gdryout = Gdryout(x) and the Gwavy curve
ceases to exist, which means that the rightmost boundary of the Gwavy
curve is its intersection with the Gdryout curve;
• If Gdryout(x) ≥ Gmist(x), which is possible at low heat fluxes and high
mass velocities, then Gdryout = Gdryout(x) and the dryout regime
disappears at this mass velocity.

Figure 18 shows four flow pattern maps calculated for R-22 at four
representative heat fluxes, where the movements of the A-D and D-M boundaries
to the left (lower x) are quite evident with increasing heat flux. Compared to the
Kattan-Thome-Favrat map, the new regimes slug (Slug), slug/stratified-wavy
(Slug+SW) and dryout (D) are now encountered. Notably, it is observed that the
dryout and mist flow regions become smaller as the heat flux decreases.
This map was developed from a database for R-22 and R-410A at 5 °C but its
prior versions covered eight other refrigerants (R-134a, R-123, R-402A, R-404A,
R-502, R-407C, R-507A and ammonia) for tube diameters from 8 to 14 mm. The
test conditions in all these experiments covered the following range of variables:
mass flow rates from 16 to 700 kgm-2s-1, vapor qualities from 1-99% and heat
fluxes from 440 to 57500 Wm-2. It is believed that the map is appropriate for
refrigerants (and other fluids with similar physical properties such as light
hydrocarbons) at low to medium pressures but not for CO2 (too high of operating
pressures for the map) nor for air-water or steam-water systems (their surface
tension and density ratios are too high with respect to the refrigerant database
noted above).
38 J.R. Thome & A. Cioncolini

Figure 18. Flow pattern maps for R-22 at 5 oC in a 13.84 mm tube evaluated at 300 kgm-2s-1 for
four heat fluxes: (a) 7.5 kWm-2, (b) 17.5 kWm-2, (c) 37.5 kWm-2, (d) 57.5 kWm-2.

7. Flow Pattern Maps for Condensation in Horizontal Tubes

The diabatic flow patterns that occur during condensation inside horizontal tubes
are similar to those for evaporation described in the previous section, but with the
following exceptions:

1. The entrance condition in a heat transfer tube usually begins with the
desuperheating of the incoming vapor that has just left the compressor, a
process that can be dry wall or wet wall depending if the wall temperature is
above or below the local saturation temperature at the local pressure. Hence,
the process tends to begin without any entrainment of liquid while for
evaporating flows the liquid bridging across the flow channel in churn and
Two-Phase Flow Pattern Maps for Macrochannels 39

intermittent flow can result in significant entrainment when these flow


structures break up.

2. During evaporation, the annular film may dry out or become entrained due to
shear forces whilst for convective condensation any such dry patch will
quickly be rewetted by a fresh condensate film and thus maintain the
integrity of the annular film. In fact, for condensation at high vapor qualities,
the flow is annular starting at the inlet vapor quality of 1.0.

3. During condensation in stratified flow regimes, the top of the tube is wetted
by the condensate film whilst in adiabatic and evaporating flows the top
perimeter is dry, a fact that blurs the distinction between annular and
stratified flows.

Therefore, the three principal flow patterns encountered by condensing flows


inside horizontal tubes are:

• Annular flow (also referred to as the shear-controlled regime in condensation


heat transfer literature);

• Stratified-wavy flow (characterized by waves on the interface of the stratified


liquid flowing along the bottom of the tube with film condensation around
the top perimeter);

• Stratified flow (no interfacial waves on the stratified liquid flowing along the
bottom of the tube with film condensation on the top perimeter that drains
downward into the stratified liquid).

The latter two regimes are also referred to as the gravity-controlled regime in
condensation heat transfer literature.
The transitions between these flow regimes can tentatively be predicted using
the Kattan et al. (1998a) or Wojtan et al. (2004) flow pattern maps for in-tube
evaporation as proposed by El Hajal et al. (2003), making the following
modifications. First, the mist flow transition is eliminated because flow in this
zone may be considered to be annular flow since a condensate film is always
formed on the cold channel wall, even if the liquid is then entrained. Secondly,
the stratified-wavy transition curve is modified by first eliminating the transition
from annular flow to stratified-wavy flow at high vapor qualities by solving for
the minimum in the stratified-wavy flow transition curve and then extending this
40 J.R. Thome & A. Cioncolini

curve as a straight horizontal line from that point to the end of the stratified
transition curve for Gstrat at a vapor quality of x = 1.0. Figure 19 illustrates this
flow pattern map for condensation of R-134a at 40 °C in a horizontal tube of 8.0
mm internal diameter. Several other methods applied to condensation are
described below.

Figure 19. Flow pattern map for in-tube condensation of R-134a at 40 °C in a horizontal tube of 8.0
mm internal diameter by El Hajal et al. (2003).

Soliman (1982) proposed a method for predicting the transition from annular
flow to stratified-wavy flow during condensation inside horizontal tubes, a
method later used by Dobson and Chato (1998) in their in-tube condensation heat
transfer model to distinguish between stratified and un-stratified flows. His
transition is based on the Froude number Frl, given for Rel ≤ 1250 as:
1.5
1.59 ⎛ 1 + 1.09 X tt0.039 ⎞ 1
Frl = 0.025Re l
⎜⎜ ⎟⎟ 0.5
(63)
⎝ X tt ⎠ Gal
and for Rel > 1250 as
1.5
1.04 ⎛ 1 + 1.09 X tt0.039 ⎞ 1
Frl = 1.26Re l
⎜⎜ ⎟⎟ 0.5
(64)
⎝ X tt ⎠ Gal
Two-Phase Flow Pattern Maps for Macrochannels 41

Soliman set the transition from annular flow to wavy flow at the value of Frl
= 7, but Dobson and Chato found that a value of 20 worked better for fitting their
condensation heat transfer correlations to their data. The superficial liquid
Reynolds number Rel used by Soliman was defined as
G(1 − x )d
Re l = (65)
µl
while his Froude number was defined as

G2
Frl = (66)
ρl2 g d
This method does not, however, give a smooth transition between these two
expressions at Rel = 1250, creating a local “jump” in the transition line when
passing through this condition. The Galileo number and Martinelli parameter
were defined and determined respectively as follows:

g ρl (ρl − ρ g )d 3
Gal = (67)
µl2
0.9 0.5 0.1
⎛1− x ⎞ ⎛ ρg ⎞ ⎛ µl ⎞
X tt = ⎜ ⎟ ⎜⎜ ⎟⎟ ⎜ ⎟ (68)
⎝ x ⎠ ⎜µ ⎟
⎝ ρl ⎠ ⎝ g ⎠

Breber et al. (1980) proposed a flow pattern map specifically designed for
condensation of pure fluids in horizontal tubes. The map, depicted in Fig. 20,
uses the Martinelli parameter X and the gas Froude number Frg as coordinates:
1/ 2
⎡ (dp / dz )l ⎤ Gg
X =⎢ ⎥ ; Frg = (69)
⎢⎣ (dp / dz )g ⎥⎦ [
ρ g (ρl − ρ g )d g 1 / 2 ]
As can be seen in Fig. 20, all transition lines are straight and either horizontal
or vertical, and this makes this map particularly easy to implement in computer
simulations. The regions between transition lines are treated as transitional flow.

Tandon et al. (1982) proposed a flow pattern map specifically designed for
condensation of pure fluids in horizontal tubes. Their map, depicted in Fig. 21,
uses the gas Froude number Frg in Eq. (69) as vertical coordinate and (1-εs)/εs as
42 J.R. Thome & A. Cioncolini

horizontal coordinate, where εs is the void fraction predicted with the Smith
(1969) expression:
−1
⎡ ⎛ ρ ⎞⎛ 1 − x ⎞⎛ ρl / ρ g + 0.4(1 − x) / x ⎞⎟⎤
ε s = ⎢1 + ⎜⎜ g ⎟⎟⎜ ⎟⎜ 0.4 + 0.6 ⎥ (70)
⎢⎣ ⎝ ρl ⎠⎝ x ⎠⎜⎝ 1 + 0.4(1 − x) / x ⎟⎠⎥

In their map, Tandon et al. (1982) label annular flow with entrained liquid
droplets as spray flow. As can be seen in Fig. 21, all transition lines are straight
and either horizontal or vertical, and this makes this map particularly easy to
implement in computer simulations.

2
10

1
10 ANNULAR BUBBLY
Froude Number

1.5
0
10 Transition Transition

0.5

-1
10 STRATIFIED INTERMITTENT

-2
1.0 1.5
10 -2 -1 0 1 2
10 10 10 10 10
Martinelli Parameter

Figure 20. Two-phase flow pattern map of Breber et al. (1980) for condensation in horizontal
tubes.

1
10 6.0 SPRAY

ANNULAR and SEMI-ANNULAR


0
10
1.0 0.5
Froude Number

-1
10 SLUG
WAVY

-2
0.01
10

PLUG

-3 0.5
10 -3 -2 -1 0 1
10 10 10 10 10
(1-Es)E-1
s

Figure 21. Two-phase flow pattern map of Tandon et al. (1982) for condensation in horizontal
tubes.
Two-Phase Flow Pattern Maps for Macrochannels 43

The flow maps of Breber et al. (1980) and Tandon et al. (1982) have been
recently extended by Cioncolini and Thome (2013) to explicitly account for
symmetric and asymmetric annular flow, similarly to what was already discussed
for the flow map of Taitel and Dukler (1976). These updated maps are depicted
in Fig. 22 and Fig. 23, respectively.

2
10

ANNULAR SYMMETRIC

1 10
10 BUBBLY

ANNULAR ASYMMETRIC
Froude Number

1.5
0
10 Transition Transition

0.5

-1
10 STRATIFIED INTERMITTENT

-2
1.0 1.5
10 -2 -1 0 1 2
10 10 10 10 10
Martinelli Parameter

Figure 22. Two-phase flow pattern map of Breber et al. (1980) for condensation in horizontal tubes
updated according to Cioncolini and Thome (2013).

1 10 SPRAY SYMMETRIC
10
SPRAY ASYMMETRIC
6.0
ANNULAR and SEMI-ANNULAR
ASYMMETRIC
0
10
1.0 0.5
Froude Number

-1
10 SLUG
WAVY

-2
0.01
10

PLUG

-3 0.5
10 -3 -2 -1 0 1
10 10 10 10 10
(1-Es)E-1
s

Figure 23. Two-phase flow pattern map of Tandon et al. (1982) for condensation in horizontal
tubes updated according to Cioncolini and Thome (2013).
44 J.R. Thome & A. Cioncolini

8. Nomenclature

A channel cross-sectional flow area (m2)


Ag channel cross-sectional area occupied by vapor (m2)
Al channel cross-sectional area occupied by liquid (m2)
Agd non-dimensional area occupied by vapor (-)
Ald non-dimensional area occupied by liquid (-)
d channel internal diameter (m)
fk friction factor of phase k (-)
F1 correction factor 1 (-)
F2 correction factor 2 (-)
Frg vapor Froude number (-)
Frl liquid Froude number (-)
g acceleration of gravity (ms-2)
G mass flux (kgm-2s-1)
Gal Galileo number (-)
Gbubbly mass flux of bubbly flow pattern transition (kgm-2s-1)
Gdryout mass flux of onset of dryout (kgm-2s-1)
Gg mass flux of gas or vapor phase flowing alone in channel (kgm-2s-1)
Gl mass flux of liquid phase flowing alone in channel (kgm-2s-1)
Gmin mass flux of minimum in mist flow pattern transition (kgm-2s-1)
Gmist mass flux of mist flow pattern transition (kgm-2s-1)
Gstrat mass flux of stratified flow pattern transition (kgm-2s-1)
Gwavy mass flux of wavy flow pattern transition (kgm-2s-1)
h liquid height in channel from bottom (m)
hld non-dimensional liquid height in channel from bottom (-)
igl latent heat of vaporization (Jkg-1)
K flow pattern map parameter (-)
dp/dz pressure drop gradient (Pam-1)
q channel wall heat flux (Wm-2)
qDNB departure from nucleate boiling heat flux (Wm-2)
Pg vapor phase perimeter of channel (m)
Pgd non-dimensional vapor phase perimeter of channel (-)
Pi phase interface length (m)
Pid non-dimensional phase interface length (-)
Pl liquid phase perimeter of channel (m)
Pld non-dimensional liquid phase perimeter of channel (-)
Reg vapor Reynolds number (-)
Rek Reynolds number of phase k (-)
Rel liquid Reynolds number (-)
ta average liquid film thickness of annular flow (m)
tb liquid film thickness at bottom of channel of annular flow (m)
T flow pattern map parameter (-)
Two-Phase Flow Pattern Maps for Macrochannels 45

Weg vapor Weber number (-)


Wel liquid Weber number (-)
x vapor quality (-)
xde vapor quality at complete dryout (-)
xdi vapor quality at onset of dryout (-)
xIA vapor quality at intermittent to annular flow transition (-)
X flow pattern map coordinate (-)
X, Xtt Martinelli parameter (tt refers to both phases turbulent) (-)
Y flow pattern map coordinate (-)
θdry dry angle around upper perimeter of channel (rad)
θstrat stratified angle around upper channel perimeter to stratified liquid level (rad)
λ flow pattern map parameter (-)
µg gas or vapor viscosity (kgm-1s-1)
µl liquid viscosity (kgm-1s-1)
µwater viscosity of water (kgm-1s-1)
ε cross sectional void fraction (-)
εS cross sectional void fraction according to Smith expression (-)
1-εS liquid holdup (-)
ξPh friction factor (-)
ρair density of air (kgm-3)
ρg vapor density (kgm-3)
ρl liquid density (kgm-3)
ρwater density of water (kgm-3)
σ surface tension (Nm-1)
Ψ flow pattern map parameter (-)

9. References

Baker, O. (1954). Simultaneous flow of oil and gas. Oil and Gas J., July issue, pp. 185-195.
Bandarra Filho, E.P. and Sáiz-Jabardo, J.M. (2006). Convective boiling performance of refrigerant
R-134a in herringbone and microfin copper tubes. Int. J. Refrigeration 29, pp. 81-91.
Barbieri, P.E.L., Sáiz-Jabardo, J.M. and Bandarra Filho, E.P. (2005). Nucleate and
convective boiling of refrigerants. Invited lecture by Sáiz-Jabardo at LTCM-EPFL,
Lausanne, Switzerland.
Barbieri, P.E.L., and Sáiz-Jabardo, J.M. (2006). The effect of the diameter in convective boiling of
refrigerant R-134a. Proc. 13th International Heat Transfer Conference, Sydney, Australia,
August 14-18.
Barnea, D. and Taitel, Y. (1986). Flow pattern transition in two-phase gas-liquid flows.
Encyclopedia of Fluid Mechanics 3, Gulf Publishing, pp. 403-474.
Biberg, D. (1999). An explicit approximation for the wetted angle in two-phase stratified
pipe flow. Canadian J. Chemical Engineering 77, pp. 1221-1224.
46 J.R. Thome & A. Cioncolini

Breber, G., Palen, J.W. and Taborek, J. (1980). Prediction of horizontal tubeside
condensation of pure components using flow regime criteria. ASME J. Heat
Transfer 102, pp. 471-476.
Bukasa, J.P., Liebenberg, L. and Meyer, J.P. (2004). Heat transfer performance during
condensation inside spiraled micro-fin tubes. ASME J. Heat Transfer 126, pp. 321-
328.
Cavallini, A., Censi, G., Del Col, D., Doretti, L., Longo, G.A. and Rossetto, L. (2002).
In-tube condensation of halogenated refrigerants. ASHRAE Trans. 108, paper 4507.
Cioncolini, A. and Thome, J.R. (2013). Liquid film circumferential asymmetry prediction
in horizontal annular two-phase flow. Int. J. Multiphase Flow 51, pp. 44-54.
Collier, J.G. and Thome, J.R. (1994). Convective Boiling and Condensation, 3rd Edition.
Oxford University Press, Oxford.
Dobson, M.K. and Chato, J.C. (1998). Condensation in smooth horizontal tubes. ASME J.
Heat Transfer 120, pp. 193-213.
El Hajal, J., Thome, J.R. and Cavallini, A. (2003). Condensation in horizontal tubes, Part
1: Two-phase flow pattern map, Int. J. Heat Mass Transfer 46, pp. 3349-3363.
Fair, J.R. (1960). What you need to know to design thermosyphon reboilers. Petroleum
Refiner 39(2), pp. 105-124.
Hewitt, G.F. and Roberts, D.N. (1969). Studies of two-phase flow patterns by
simultaneous X-ray and flash photography. AERE-M 2159, HMSO.
Kattan, N., Thome, J.R. and Favrat, D. (1998a). Flow boiling in horizontal tubes. Part 1:
development of a diabatic two-phase flow pattern map. ASME J. Heat Transfer 120, pp. 140-
147.
Kattan, N., Thome, J.R. and Favrat, D. (1998b). Flow boiling in horizontal tubes. Part 2: new heat
transfer data for five refrigerants. ASME J. Heat Transfer 120, pp. 148-155.
Kattan, N., Thome, J.R. and Favrat, D. (1998c). Flow boiling in horizontal tubes. Part 3:
development of a new heat transfer model based on flow patterns. ASME J. Heat Transfer
120, pp. 156-165.
Kutateladze, S.S. (1948). On the transition to film boiling under natural convection.
Kotloturbostroenie 3, pp. 152-158.
Lavin, J.G. and Young, E.H. (1965). Heat transfer to evaporating refrigerants in two-
phase flow. AIChE J. 11, pp. 1124-1132.
Liebenberg, L., Thome, J.R. and Meyer, J.P. (2005). Flow visualization and flow pattern
identification with power spectral density distributions of pressure traces during
refrigerant condensation in smooth and micro-fin tubes. ASME J. Heat Transfer 27,
pp. 209-220.
Meyer, J.P. and Liebenberg, L. (2006). In search of objective heat transfer and pressure
drop models. Proc. 13th International Heat Transfer Conference, Sydney, Australia,
August 14-18.
Mori, H., Yoshida, S., Ohishi, K. and Kokimoto, Y. (2000). Dryout quality and post-
dryout heat transfer coefficient in horizontal evaporator tubes. Proc. of 3rd
European Thermal Sciences Conference, pp. 839-844.
Olivier, J.A., Liebenberg, L., Kedzierski, M.A. and Meyer, J.P. (2004). Pressure drop
during refrigerant condensation inside horizontal smooth, helical micro-fin and
herringbone micro-fin tubes. ASME J. Heat Transfer 126, pp. 687-696.
Two-Phase Flow Pattern Maps for Macrochannels 47

Olivier, J.A., Liebenberg, L., Thome, J.R. and Meyer, J.P. (2007). Heat transfer, pressure
drop, and flow pattern recognition during condensation inside smooth, helical
micro-fin and herringbone tubes. Int. J. Refrigeration 30, pp. 609-623.
Rouhani, S.Z. and Axelsson, E. (1970). Calculation of void volume fraction in the
subcooled and quality boiling regions. Int. J. Heat Mass Transfer 13, pp. 383‒393.
Silva Lima, R.J. (2011). Experimental and visual study on flow patterns and pressure
drops in U-tubes. Ph.D. Thesis, Swiss Federal Institute of Technology-EPFL.
Smith, S.L. (1969). Void fraction in two-phase flow: a correlation based upon an equal
velocity head model. Proc. Inst. Mech. Eng. Lond. 184, pp. 647–657, Part I.
Soliman, H.M. (1982). On the annular-to-wavy flow pattern transition during
condensation inside horizontal tubes. Canadian J. Chem. Eng. 60, pp. 475-481.
Steiner, D. (1993). VDI-Wärmeatlas (VDI Heat Atlas), Verein Deutscher Ingenieure,
VDI-Gesellschaft Verfahrenstechnik und Chemieingenieurwesen (GCV),
Düsseldorf, Chapter Hbb.
Taitel, Y. and Dukler, A.E. (1976). A model for predicting flow regime transitions in
horizontal and near horizontal gas-liquid flow, AIChE J. 22, pp. 47-55.
Taitel, Y. (1990). Flow pattern transition in two-phase flow. Proc. 9th International Heat
Transfer Conference, Jerusalem 1, pp. 237-254.
Tandon, T.N., Varma, H.K. and Gupta, C.P. (1982). A new flow regime map for
condensation inside horizontal tubes. ASME J. Heat Transfer, 104, pp. 763-768.
Thome, J.R. and El Hajal, J. (2003). Two-phase flow pattern map for evaporation in
horizontal tubes: latest version. Heat Transfer Engineering 24, pp. 3-10.
Ursenbacher, T., Wojtan, L. and Thome, J.R. (2004). Dynamic void fractions in stratified
types of flow, Part I: New optical measurement technique and dry angle
measurements. Int. J. Multiphase Flow 30, pp. 107-124.
Wojtan, L., Ursenbacher, T. and Thome, J.R. (2004). Dynamic void fractions in stratified
types of flow, Part II: Measurements for R-22 and R-410a. Int. J. Multiphase Flow
30, pp. 125-137.
Wojtan, L., Ursenbacher, T. and Thome, J.R. (2005). Investigation of flow boiling in
horizontal tubes: Part I – A new diabatic two-phase flow pattern map, Int. J. Heat
Mass Transfer 48, pp. 2955-2969.
Zürcher, O. Thome, J.R. and Favrat, D. (1997). Prediction of two-phase flow patterns for
evaporation of refrigerant R-407C inside horizontal tubes. ECI Convective Flow
and Pool Boiling Conference, Irsee, Germany, May 18-23, paper IX-1.
Zürcher, O.,Thome, J.R. and Favrat, D. (1999). Evaporation of ammonia in a smooth horizontal
tube: heat transfer measurements and predictions. ASME J. Heat Transfer 121, pp. 89-101.
Zürcher, O., Favrat, D. and Thome, J.R. (2002). Development of a diabatic two-phase
flow pattern map for horizontal flow boiling. Int. J. Heat Mass Transfer 45, pp.
291-303.

View publication stats

You might also like