Vous êtes sur la page 1sur 24

VOL. 9, NO.

1 HVAC&R RESEARCH JANUARY 2003


55
Accurate Charge Inventory Modeling
for Unitary Air Conditioners
Todd M. Harms, Ph.D. Eckhard A. Groll, Ph.D. James E. Braun, Ph.D.
Member ASHRAE Member ASHRAE
Charge inventory, the accounting of the distributed fluid mass in closed systems, is necessary to
predict system performance at different environmental conditions for a given charge. Public
domain simulation models used to predict unitary equipment performance are currently unable
to accurately determine charge inventory. Sources of error in these models include incomplete
internal volume accounting, neglected refrigerant-oil diffusion effects, and void fraction model-
ing assumptions. The results presented here suggest that the most important charge inventory
issue is proper void fraction determination. The Baroczy void fraction correlation gave the best
agreement with measured data for three unitary air conditioners. As the condenser holds by far
the largest percentage of the total charge, accurate prediction of heat transfer and pressure
drop in the condenser was found to be necessary for charge inventory modeling.
INTRODUCTION
Charge inventory strongly influences the overall performance of unitary air conditioners. For
modeling purposes, if the charge is not specified, then another system parameter, such as con-
denser subcooling, must be specified to establish a unique operating condition. In general, the
charge necessary to achieve a particular subcooling varies with operating conditions, and there-
fore this approach is only appropriate for a single design-point calculation. For this reason, sys-
tem models typically incorporate charge inventory calculations.
Two commonly used system models are HPSIM (Domanski and Didion 1983) and PUREZ
(Rice and Jackson 1994). During the validation process, these models were found to underpre-
dict charge inventory. Several sources of error have been identified in modeling of charge inven-
tory, including incomplete internal volume accounting, neglected refrigerant-oil diffusion
effects, and void fraction modeling assumptions (Damasceno et al. 1991; Marques and Melo
1993). Perhaps the most challenging of these is void fraction determination.
At a given cross section of duct, the void fraction is defined as the fraction of area occupied
by vapor. Assuming that the quality of the refrigerant can be determined using conservation
equations, a void fraction model is needed to relate two-phase density to quality. Rice (1987)
presented a comprehensive review of the available void fraction models. The void fraction cor-
relations of Baroczy (1965), Hughmark (1962), Premoli et al. (1971), and Tandon et al. (1985)
were recommended, because they yield the highest charge predictions for condensers and the
best overall agreement with experimental data. Rice stated that there are insufficient data to rec-
ommend one over the others; the Hughmark method may overpredict charge in the condenser,
yet still yield good agreement with the total charge by way of error cancellation with respect to
unaccounted charge elsewhere in the system.
Todd M. Harms is a consulting engineer in Plymouth, Minnesota. Eckhard A. Groll is an associate professor of me-
chanical engineering and James E. Braun is a professor of mechanical engineering at Purdue University, Ray W. Herrick
Laboratories, West Lafayette, Indiana.
Harms474.fm Page 55 Monday, December 16, 2002 8:10 AM
56 HVAC&R RESEARCH
Other studies have compared predictive models with experiments. Damasceno et al. (1991)
used HPSIM to compare predicted and measured capacity for various charge levels in a residen-
tial air-to-air heat pump. They found that they needed to modify HPSIM to include the Hugh-
mark (1962) correlation to get acceptable results. They did not consider oil diffusion. Marques
and Melo (1993) compared the predicted and measured charge inventory of a room air condi-
tioner using HPSIM. They also found that the Hughmark correlation provided the best results.
Furthermore, they found that it was necessary to add an oil diffusion calculation to HPSIM.
LeRoy et al. (2000) used PUREZ to compare the predicted and measured charge for seven uni-
tary heat pumps. They found that the model consistently underpredicted charge, and that the
Hughmark model gave the best agreement with the data. PUREZ does not consider oil diffusion.
In an experimental study, Mulroy and Didion (1985) measured the charge inventory in five
sections (the compressor and accumulator, the liquid line, the vapor line, the condenser, and the
evaporator) of a nominally 3 ton (10.6 kW) split-system air conditioner equipped with a capil-
lary tube expansion device. To isolate the various sections of the system, five pneumatic valves
were simultaneously closed during system operation. The sections were then detached and
weighed. The system was only tested at one charge level. At steady-state conditions, 83% of the
charge resided in the condenser and the liquid line.
The literature suggests that several aspects of charge inventory modeling need investigation,
and that the total charge inventory is routinely underpredicted. If the charge inventory has been
underpredicted in previous studies and void fraction correlations were selected to compensate
for unaccounted charge, then the most accurate void fraction correlation available has yet to be
identified. Because not all researchers include the same system sources to account for charge,
the relative importance of the various charge locations seems to be in doubt.
This paper presents the results of a study that investigated the effects of various void fraction
correlations on refrigerant charge predictions. To reduce propagation of modeling errors, very
detailed component models were developed and used separately to predict refrigerant charge
using measured refrigerant and air-side inlet conditions. Total charge inventory was predicted
by summing the contributions from the component models. The heat exchanger model predic-
tions of heat transfer and pressure drop were validated using the measurements. In addition, all
the volumes within the system were carefully accounted for and refrigerant absorbed within oil
in the compressor sump was included. Total charge inventory predictions are compared for
using different void fraction correlations in the heat exchanger models. The relative magnitude
of the charge inventory predictions is presented for each of the components to study charge on
the basis of system location.
EXPERIMENTAL APPARATUS AND PROCEDURE
The experimental work used three unitary air conditioners; their physical characteristics are
given in Table 1. Between the three systems, there was some variation in the nominal capacity,
refrigerant, compressor type, and compressor oil. All three systems were designed for R-22.
However, the refrigerant in the 2.5 ton (8.8 kW) split system was replaced with R-407C, which
required that the compressor oil also be changed. All three systems used a thermal expansion
valve for the expansion device. Each heat exchanger had multiple, parallel-flow circuits to man-
age pressure drop, and most had more than one row of tubes. All of the heat exchangers used
standard microfin tubes, except two of the condensers. The condenser tubing in the 2.5 ton (8.8
kW) split system had longer, less sloped microfins and lacked the helical twist seen in the other
heat exchangers. The condenser tubing in the 7.5 ton (26.4 kW) split system lacked microfins.
The listed microfin parameters are illustrated in Figure 1.
A schematic of the test setup, including the location and type of each sensor, is given in
Figure 2. All temperature measurements were obtained using Type T (copper-constantan)
Harms474.fm Page 56 Monday, December 16, 2002 8:10 AM
VOLUME 9, NUMBER 1, JANUARY 2003 57
thermocouples. The thermocouples used to measure air temperatures and surface temperatures
were made from the same spool of wire. Surface thermocouples were soldered to the surface of
interest to minimize contact resistance. Refrigerant temperatures were measured using sheathed
probe thermocouples. The accuracy of all of these thermocouples was increased to 0.2 K by
calibration.
A thermocouple probe was installed in the refrigerant flow at each of the eight state points
except 3 and 7. The thermodynamic states at points 2 and 3 were assumed to be equivalent
because the discharge line was relatively short. Thermocouples were mounted on the surface of
each distributor line to provide an average temperature for point 7. A three-by-three grid of ther-
mocouples was installed at the inlet and exit of each heat exchanger. An array of thermocouples
was mounted from top to bottom on the compressor shell. Pressure transducers were used to
measure the refrigerant pressure at each of the state points except 3 and 7. These transducers
were calibrated to an accuracy of 20 kPa by way of a dead-weight tester.
A chilled mirror sensor was used to measure the dew point (accurate to 0.2 K) of the air at the
evaporator inlet and outlet. An air valve was used to alter the sensing location. The mass flow
rate of condensed water exiting the evaporator was measured using a stopwatch and a digital
scale accurate to 0.01 g. The mass flow rate of refrigerant was measured using a Coriolis-type
sensor, installed in the liquid line with sight glasses on either side. The flow meter was cali-
brated to an accuracy of 0.27 g/s by measuring the average flow rate over a range of values with
a stopwatch and a digital scale. The mass of refrigerant added to the system was measured with
Table 1. Physical Characteristics of Three Experimental Systems
2.5 Ton (8.8 kW)
Split System
5 Ton (17.6 kW)
Rooftop Unit
7.5 Ton (26.4 kW)
Split System
Refrigerant R-407C R-22 R-22
Compressor Reciprocating Scroll Reciprocating
Compressor oil Polyol ester oil Polyol ester oil Blended mineral oil
Expansion device Thermal expansion valve Thermal expansion valve Thermal expansion valve
Condenser Evaporator Condenser Evaporator Condenser Evaporator
Circuits 2 6 6 8 7 10
Rows 1 3 2 4 2 3
D
i
7.62 mm 8.99 mm 8.99 mm 8.99 mm 9.19 mm 8.99 mm
No. of microfins 60 60 60 60 60
Apex angle 30 50 50 50 50
Helix angle 0 18 18 18 18
Microfin height 0.33 mm 0.20 mm 0.20 mm 0.20 mm 0.20 mm
Figure 1. Schematics of Microfin Tube
Harms474.fm Page 57 Monday, December 16, 2002 8:10 AM
58 HVAC&R RESEARCH
a digital scale accurate to 0.005 kg. The indoor blower power and the outdoor fan power were
measured with watt transducers accurate to 10 W. Atmospheric pressure was measured using a
mercury barometer accurate to 0.03 kPa.
The experiments were carried out in ASHRAE standard psychrometric chambers (ASHRAE
1988) to set and maintain specific environmental conditions. The indoor room was equipped with
an air measurement box as specified by the standard. The evaporator air circuit of each system
was connected in series to this box to measure the flow rate of air. The measurement was based
on Bernoullis equation applied across several parallel, long-radius nozzles. Two differential
pressure transducers were installed in the box to measure the pressure drop across the nozzles and
the nozzle pressure, respectively. Both transducers were calibrated against a Hooke gage
manometer accurate to 0.01 mm H
2
O (0.1 Pa). Three resistance temperature devices (RTDs),
specified to be 0.1% accurate, measured the air temperature at the inlet to the nozzles. For each
system, the volumetric flow rate of air through the condenser was given by the manufacturer and
was taken to be constant.
The input parameters varied during testing were the environmental conditions and the refrig-
erant charge. The range of tested environmental conditions is listed in Table 2. Test conditions
Table 2. Environmental Conditions
Test Condition T
outdoor
, C T
indoor
, C T
dew point
, C
A 35.0 26.7 15.8
B 27.8 26.7 15.8
C 27.8 26.7 <3.1
HT 48.9 26.7 15.8
Figure 2. Schematic of Test Setup
Harms474.fm Page 58 Monday, December 16, 2002 8:10 AM
VOLUME 9, NUMBER 1, JANUARY 2003 59
A and B represent standard conditions, and only differ in the outdoor air temperature; the low
dew point specified for test condition C was designed to keep the evaporator coil dry (ASHRAE
1995). The HT condition was intended to simulate the high outdoor air temperature of a desert
environment.
Starting from an evacuated state, the system was initially charged to achieve nominal sub-
cooling. Data were recorded once steady-state conditions were achieved in the chambers and in
the system. Refrigerant was metered into the system to achieve increased charge inventory at
discrete levels. At each level of charge inventory, the system was allowed to reach steady state
before data were recorded. This process was repeated for each set of environmental conditions.
The three systems were tested in succession, allowing the same instrumentation to be used on
each system. The test matrix for the three systems is given in Table 3. All three units were tested
at test condition A over a range of charge levels. The 5 ton (17.6 kW) rooftop unit was also
tested at test conditions B, C, and HT, for a range of charge levels. Subcooling ranged from just
above zero to almost 20 K. In contrast, superheat was maintained in a much smaller range for a
given set of charge-varied tests, as expected for systems with a thermal expansion valve. The
mass flux ranged from 250 to 400 kg/(m
2
s) for the condensers and from 100 to 250 kg/(m
2
s)
for the evaporators.
COMPONENT MODELING
To predict the refrigerant charge inside the experimental systems, individual component mod-
els were used with measured state points as inputs. The total charge inventory was obtained by
summing the contributions of the component models. Because of the complexity of two-phase
flow, the heat exchanger models were the primary focus of this work. The other component
models account for the mass of refrigerant in the line sets and the compressor. The line sets were
modeled as adiabatic and frictionless. The compressor model accounts for the volume of the
shell and the refrigerant dissolved in the oil.
Heat Exchanger Modeling
The heat exchangers consisted of one to four rows of tubes. For modeling purposes, each tube
within a row was sectioned into several tube elements. Mass, energy, and momentum balances
were applied to the air and refrigerant in each element to obtain heat transfer rates and outlet
states for given inlet conditions. Each heat exchanger was sectioned to give approximately a
0.01 change in quality across two-phase elements. Because segmentation was fairly refined, the
properties were assumed to be constant within an element. The properties were updated at suc-
cessive elements. Saturation pressure drops and the corresponding saturation temperature drops
were considered for the refrigerant. If the air state for a given element was not known a priori,
then an iterative solution was applied.
Table 3. Test Matrix for Systems
Exp. # Condition Charge, kg
Condenser Evaporator
T
sub
, K G, kg/(m
2
s) T
super
, K G, kg/(m
2
s)
2.5.A1-4 A 2.3 to 3.1 1.4 to 18.1 381 to 398 5.1 to 5.4 95 to 100
5.A1-5.A7 A 4.3 to 7.2 2.7 to 14.3 274 to 277 7.3 to 9.2 205 to 208
5.B1-5.B7 B 3.9 to 6.4 0.3 to 9.8 253 to 270 8.7 to 14.4 190 to 203
5.C1-5.C7 C 4.0 to 7.4 1.3 to 14.1 249 to 253 7.3 to 10.7 187 to 189
5.HT1-3 HT 4.3 to 6.1 1.0 to 10.3 275 to 278 6.3 to 7.9 206 to 208
7.5.A1-3 A 6.2 to 7.6 1.5 to 12.1 352 to 358 5.9 to7.6 241 to 245
Harms474.fm Page 59 Monday, December 16, 2002 8:10 AM
60 HVAC&R RESEARCH
Condenser Elements. In the condenser, one-dimensional heat transfer was assumed, so that
(1)
where C
min
and were determined according to the NTU- method (Incropera and DeWitt
1996):
(2)
The form of the effectiveness depends on the arrangement of the heat exchanger. For single-phase
flow in a cross-flow heat exchanger where one fluid (air) is unmixed and the other fluid (refrig-
erant) is mixed, the effectiveness is determined as
(3)
where
(4)
For two-phase flow in any heat exchanger element where the saturation temperature is constant,
the effectiveness is
(5)
The number of transfer units necessary to evaluate effectiveness depends on the overall conduc-
tance according to
(6)
where
(7)
Q

cond
C
min
T
ref
T
air
( ) =
C
min
min
m

ref
c
p ref ,
m

air
c
p air ,

1
C
r
------ 1 exp C
r
1 exp NTU ( ) [ ] { } ( ) m

ref
c
p ref ,
m

air
c
p air ,
>
1 exp
1
C
----

r
1 exp C
r
NTU ( ) [ ]
)
`

m

air
c
p air ,
m

ref
c
p ref ,
>

=
C
r
min
m

ref
c
p ref ,
m

air
c
p air ,

air
c
p air ,
m

ref
c
p ref ,

=
1 exp NTU ( ) =
NTU
UA
C
min
------------
=
1
UA
--------
1
h
i
A
i
----------
ln
D
o
D
i
------
2kL
-------------
1

o
h
o
A
o
------------------
+ + =
Harms474.fm Page 60 Monday, December 16, 2002 8:10 AM
VOLUME 9, NUMBER 1, JANUARY 2003 61
The overall fin efficiency is defined as
(8)
where
(9)
and
(10)
The effective fin length L
f
for a plate fin heat exchanger was determined using the empirical
method presented by Schmidt (1949). The inside heat transfer coefficient in the two-phase
region was calculated based on the work of Cavallini and Zecchin (1974) for smooth tubes and
Cavallini et al. (2000) for microfin tubes; the Gnielinski (1976) correlation was used in the
single-phase region. The outside heat transfer coefficient was obtained from the equipment
manufacturer.
For the refrigerant side, an energy balance on a tube element provides the outlet refrigerant
enthalpy:
(11)
For the air side, an energy balance on a tube element gives the outlet air temperature:
(12)
Evaporator Elements. For tube elements of the evaporator where moisture does not con-
dense, the model is identical to the condenser model, so that
(13)
where C
min
and were again determined according to the NTU- method. Calculation of the
inside and outside heat transfer coefficients follows that of the condenser except that the
two-phase inside heat transfer coefficient is based on the work of Chen (1966). The surface area
enhancement of the internal microfins was taken into account. However, no additional enhance-
ment to the heat transfer coefficient was considered.
If the coil surface temperature of a given element is below the dew point of the air, then a wet
coil analysis was used based on Braun et al. (1989):
(14)

o
1
A
f
A
t
----- 1
f
( ) =

f
tanh m
f
L
f
( )
m
f
L
f
-----------------------------
=
m
f
2 w
t
t
c
+ ( )h
o
k
f
w
t
t
c
------------------------------
1 2
=
Q

cond
m

ref
h
in
h
out
( ) =
Q

cond
m

air
c
p air ,
T
out
T
in
( ) =
Q

evap
C
min
T
air
T
ref
( ) =
T
s
T
ref
Q

evap 1
h
i
A
i
----------
ln
D
o
D
i
------
2kL
------------- +
+ =
Harms474.fm Page 61 Monday, December 16, 2002 8:10 AM
62 HVAC&R RESEARCH
The wet analysis is a mechanistic model of the combined heat and mass transfer. The solution is
cast in the form of the NTU- method, using enthalpy differences instead of temperature differ-
ences:
(15)
Again, the form of the effectiveness depends on the phase of the refrigerant, where, for single-
phase flow,
(16)
where
(17)
and where, for two-phase flow,
(18)
The number of transfer units for a wet coil is given as
(19)
where
(20)
Casting the equation for conduction in a fin in terms of enthalpy differences results in several
unique quantities. The overall fin efficiency for a wet coil
o,w
is analogous to the usual form of
the overall fin efficiency
o
, but is also a function of the saturation specific heat c
s
and the spe-
cific heat of the air-water mixture c
p,m
. The saturation specific heat is defined as the derivative
with respect to temperature of the saturated air enthalpy evaluated at the refrigerant temperature:
(21)
Q

evap

w
m

air
h
a i ,
h
r s ,
( ) =

w
1
C
r w ,
---------- 1 exp C
r w ,
1 exp NTU
w
( ) [ ] { } ( ) m

ref
c
p ref ,
m

air
c
s
>
1 exp
1
C
r w ,
----------
1 exp C
r w ,
NTU
w
( ) [ ]
)
`

m

air
c
s
m

ref
c
p ref ,
>

=
C
r w ,
min
m

ref
c
p ref ,
m

air
c
s

air
c
s
m

ref
c
p ref ,

w
1 exp NTU
w
( ) =
NTU
w
UA
w
m

air
-----------
=
1
UA
w
-----------
c
s
h
i
A
i
----------
c
s
ln
D
o
D
i
------
2kL
-------------
c
p m ,

o w ,
h
o
A
o
------------------------
+ + =
c
s
dh
r s ,
dT
------------
=
Harms474.fm Page 62 Monday, December 16, 2002 8:10 AM
VOLUME 9, NUMBER 1, JANUARY 2003 63
The overall wet fin efficiency is defined as
(22)
where
(23)
and
(24)
The effective fin length L
f
is calculated as in the condenser.
For the refrigerant side, an energy balance on a tube element provides the outlet refrigerant
enthalpy:
(25)
For the air side, an energy balance on a dry tube element gives the outlet air temperature:
(26)
The air-side energy balance is more complicated when moisture removal occurs. An energy bal-
ance on a wet tube element gives the outlet air enthalpy:
(27)
One more thermodynamic property is needed to uniquely determine the state of the wet air at the
outlet of the tube element. First, the saturated air enthalpy at the tube surface is determined as
follows:
(28)
where
(29)
Next, the outlet air temperature at saturated conditions is determined from psychrometric prop-
erty routines, and is a function of the saturation air enthalpy at the tube surface and the baromet-
ric air pressure:
(30)
A rate of mass transfer analysis that is analogous to the combined heat and mass transfer rate
analysis is the next logical step. However, that would require a mass transfer fin efficiency,

o w ,
1
A
f
A
t
----- 1
f w ,
( ) =

f w ,
tanh m
f w ,
L
f
( )
m
f w ,
L
f
----------------------------------
=
m
f w ,
2 w
t
t
c
+ ( )h
o w ,
c
s
k
f
w
t
t
c
c
p m ,
-----------------------------------------
1 2
=
Q

evap
m

ref
h
out
h
in
( ) =
Q

evap
m

air
c
p m ,
T
in
T
out
( ) =
Q

evap
m

air
h
air in ,
h
air out ,
( ) =
h
s s o , ,
h
air in ,
h
air in ,
h
air out ,

1 exp NTU
o w ,
( )
---------------------------------------------
=
NTU
o w ,

o w ,
h
o w ,
A
o
m

air
c
p m ,
-----------------------------
=
T
s o ,
func h
s s o , ,
P
air
, ( ) =
Harms474.fm Page 63 Monday, December 16, 2002 8:10 AM
64 HVAC&R RESEARCH
which has yet to be defined. Instead, an analysis of the rate of sensible heat transfer from the sur-
face of the tube to the air was used to approximate the outlet air temperature:
(31)
where
(32)
Braun et al. (1989) evaluated the overall accuracy of this method for use in modeling cooling
coils. For a global heat exchanger analysis, the effectiveness was found to be within 5% of the
numerical solutions of the governing conservation equations. This method is expected to per-
form even better when applied to a small element.
Pressure Drop. Pressure drop was calculated in the same manner for both heat exchangers
(Carey 1992):
(33)
where
(34)
and
(35)
The two-phase frictional pressure drop multiplier was calculated based on the work of Lockhart
and Martinelli (1949):
(36)
where the Lockhart-Martinelli parameter is defined as
(37)
when the refrigerant is single-phase the pressure drop is given as
(38)
For both single- and two-phase flow, the friction factor was calculated according to Haaland
(1983). For this correlation, internal microfins were treated as roughening elements.
T
o
T
s o ,
T
in
T
s o ,
( )exp NTU
o
( ) + =
NTU
o

o
h
o
A
o
m

air
c
p m ,
-----------------------
=
P P
fr
P
acc
+ =
P
fr

Gx
g
( )
2
2
g
--------------------
4f
g
L
D
i
----------- =
P
acc
G
2
x
out
2

out
-----------------
1 x
out
( )
2

f
1
out
( )
-------------------------------
+ G
2
x
in
2

in
--------------
1 x
out
( )
2

f
1
in
( )
----------------------------
+ =

g
2
1 20X
tt
X
tt
2
+ + =
X
tt
1 x
x
-----------
\
| |
0.9
g

f
-----
\
| |
0.5
f

g
------
\
| |
0.1
=
P
G
2
2
------
4f L
D
i
--------- =
Harms474.fm Page 64 Monday, December 16, 2002 8:10 AM
VOLUME 9, NUMBER 1, JANUARY 2003 65
A typical method of managing pressure drop in heat exchangers is to split the flow into paral-
lel circuits. A rigorous method of determining the mass flow rate in each circuit would be to per-
form a momentum balance on the system. Because these circuits are typically the same length, a
simpler approach was used where the mass flow rate through each circuit was assumed to be the
same. This method is not exact, but does provide accurate results (Harms 2002).
Charge Inventory
The mass of refrigerant in a given volume was determined as
(39)
The length and diameter of the tubing were provided by the manufacturer or were directly measured.
For each compressor, the volume of the shell occupied by vapor was determined from the manufac-
turers. A temperature measurement on the upper part of the compressor shell was used to determine
the density of the vapor. If two-phase flow was encountered, then an apparent density was used:
(40)
where is the void fraction. Besides in the heat exchangers, two-phase flow was encountered in
the line set between the thermal expansion valve and the evaporator.
The mass of refrigerant dissolved in the compressor oil was a function of the solubility of
refrigerant-oil mixture:
(41)
Solubility data were obtained from the compressor manufacturers as a function of the tempera-
ture and pressure of the mixture. A temperature measured at the base of the compressor shell
was used as the mixture temperature. The mixture pressure was taken to be equal to the suction
pressure.
Void Fraction Correlations
Flow visualization experiments have shown that two-phase flow can exhibit a wide range of
flow regimes. In general, void fraction models are based on a particular flow regime, which can
be predicted using measurable quantities with a flow regime map. Taitel and Dukler (1976) pro-
posed the flow regime map seen in Figure 3 for horizontal, two-phase flow using a mechanistic
approach to flow regime transitions. The wavy, annular, and intermittent flow regimes are indi-
cated, along with the related flow regime transitions. For wavy flow, the respective phases are
stratified because of gravity, and are separated by a wavy interface. Annular flow consists of a
vapor core surrounded by a liquid annulus. Wavy and annular flow can both be defined as sepa-
rated flow regimes. Intermittent flow encompasses slug flow and plug flow, and is characterized
by large variations in the void fraction at a given cross section. The transitions are given in terms
of the Lockhart-Martinelli parameter and a modified form of the Froude number:
(42)
where
(43)
m V =

g

f
1 ( ) + =
m
dissolved
m
oil

1
----------- =
Fr
TD

f

g

-----------------
\
| |
0.5
Fr =
Fr
xG

g
gD
i
( )
0.5
--------------------------
=
Harms474.fm Page 65 Monday, December 16, 2002 8:10 AM
66 HVAC&R RESEARCH
Simulated transition parameters are plotted in Figure 3 for conditions typical for horizontal
heat exchangers to illustrate their physical relationship to measured quantities, such as quality
and mass flux. The parameters were determined for R-22 and a tube diameter of 9 mm. The
mass flux was set at the extremes for the present experiment, 100 and 400 kg/(m
2
s). The pres-
sure was set to 650 and 2000 kPa, to simulate conditions in an evaporator and in a condenser,
respectively. The quality ranged from 0.01 to 0.99. The results show that traversing the
wavy-flow transition boundary occurs almost exclusively by varying the mass flux. Conversely,
the transition from intermittent flow to annular is, by definition, only a function of the Lock-
hart-Martinelli parameter, which is in turn a strong function of quality. Notably, all three
regimes are important when considering unitary air conditioners. The transition to bubbly flow
from intermittent is not shown in Figure 3, but occurs at a mass flux of around 1600 kg/(m
2
s),
which is out of the range of typical air conditioners.
The void fraction models examined in this work and the flow regime for each model are listed
in Table 4. Additionally, some the characteristics of the correlations are summarized in the table.
The models are analytical, semiempirical, or empirical. Given the complexity of two-phase
flow, only very simple models are completely analytical. Here, models are deemed semiempiri-
cal if they are essentially analytical, but use empirical closure equations or fitted constants.
Two-phase empirical models often assume a functional form or a set of dimensionless parame-
ters, and use constants fit to experimental data. As expected of models that are popular for
design work, most of the correlations are explicit functions of measurable quantities.
Void fraction models for four flow regimes were considered: separated, annular, slug, and
bubbly. Separated flow is a general classification that includes annular flow and wavy flow, and
incorporates two primary assumptions: the flow is separated into continuous liquid and vapor
regions, and the velocity within each region is uniform and distinct at a given cross section. The
Figure 3. Taitel-Dukler Two-Phase Flow Regime Map for Horizontal Flow
Harms474.fm Page 66 Monday, December 16, 2002 8:10 AM
VOLUME 9, NUMBER 1, JANUARY 2003 67
ratio of these velocities is termed the slip ratio because, in general, the vapor phase has a higher
velocity:
(44)
Void fraction and mass quality are defined as follows:
(45)
and
(46)
Void fraction can be related to quality in terms of the slip ratio by applying mass conservation
and combining Equations (44) to (46):
(47)
Slug flow is a subset of intermittent flow. Slug flow and bubbly flow are somewhat similar
regimes, both being intermittent. Each void fraction model will be considered in further detail
according to flow regime.
Separated Flow. The simplest slip flow model is the homogeneous model. The slip ratio is
taken as one, meaning that the respective phases are traveling at the same mean velocity.
Zivi (1964) applied an assumption of no entropy production to generate a void fraction model.
The result gives the slip ratio as a function of the saturation densities:
(48)
This model does not account for viscous effects, which would increase the slip ratio, or entrain-
ment of liquid in the vapor, which would decrease the slip ratio.
Table 4. Void Fraction Correlations
Regime
Correlation
Type Form Notes
Homogeneous Analytical Explicit S = 1
Baroczy (1965) Separated Empirical Tabular Vertical upflow
Yashar et al. (2001) Separated Empirical Explicit Horizontal
Zivi (1964) Separated Analytical Explicit No friction
Premoli et al. (1971) Separated Empirical Explicit Vertical upflow
Tandon et al. (1985) Annular Semiempirical Explicit
Hughmark (1962) Bubbly Empirical Implicit Vertical upflow
Taitel and Barnea (1990) Slug Semiempirical Explicit
S
u
g
u
f
-----
=

A
g
A
------
=
x
m

g
m

------
=

1
1
1 x
x
-----------
\
| |

f
-----S +
------------------------------------
=
S

f
-----
\
| |
1 3
=
Harms474.fm Page 67 Monday, December 16, 2002 8:10 AM
68 HVAC&R RESEARCH
Premoli et al. (1971) correlated experimental data to an empirical separated-flow model. The
experiments were conducted in vertical ducts with adiabatic upflow conditions. The results were
optimized for predicting the apparent two-phase density. The slip ratio was correlated in terms
of several parameters:
(49)
where
(50)
and
(51)
A separated-flow model was proposed by Baroczy (1965) using void fraction data for both
air-water and nitrogen-mercury combinations. The model is given in tabular form as a function
of the Lockhart-Martinelli parameter and a saturated property term:
(52)
Yashar et al. (2001) correlated the void fraction with respect to a modified Froude number
and the Lockhart-Martinelli parameter:
(53)
where
(54)
Their correlation takes two forms, one for condensation in internally enhanced tubes and one for
all other cases:
(55)
They found that microfins tend to hold more refrigerant than smooth tubes during condensation.
Annular Flow. Tandon et al. (1985) proposed an annular flow void fraction model based on
turbulent boundary layer theory. The velocity profile in the liquid annulus was represented by
S func x

f
----- Re
L
We
L
, , ,
\
| |
=
Re
L
GD
i

f
----------
=
We
L
G
2
D
i

f
g
-------------
=
func X
n

f
-----

g
------
\
| |
,
0.2
=
1
1
Ft
-----
X
tt
+ +
\
| |
n
=
Ft
x
1 x
-----------
\
| |
0.5
Fr =
n
0.375 microfin condensation
0.321 all other cases

=
Harms474.fm Page 68 Monday, December 16, 2002 8:10 AM
VOLUME 9, NUMBER 1, JANUARY 2003 69
the von Karman velocity profile, and the Lockhart-Martinelli approach was used to calculate the
two-phase pressure drop. The result is as follows:
(56)
Slug Flow. Taitel and Barnea (1990) modeled slug flow as slug units consisting of two
regions: a primarily liquid slug region and a bubble region with stratified liquid. If the slug is
assumed to be entirely liquid, then the void fraction equation reduces to
(57)
Bubbly Flow. Hughmark (1962) built on the work of Bankoff (1960), who developed a void
fraction model for bubbly flow. A power law distribution was assumed for the velocity profile of
the bubbles. The resulting correlation is as follows:
(58)
Hughmark correlated the parameter K to data from vertical upflow experiments. For conditions
typical to air conditioners, K ranges from 0.69 to 0.97; it increases with increasing quality and is
weakly dependent on mass flux.
Figure 4 compares slip ratio predictions for the seven void fraction correlations presented
here. Predictions were generated for R-22 at 2 MPa, in a 9 mm diameter tube, at a mass flux of
250 kg/(m
2
s), typical conditions for a condenser in a unitary air conditioner. For the Hughmark
and the Taitel and Barnea correlations, the slip ratio is only an effective value because these cor-
relations are not based on separated flow regimes. An effective slip ratio was determined using
Equation (47). The quality at which intermittent to annular transition takes place is indicated in
the figure. The deficiency of the Hughmark and the Taitel and Barnea models beyond the inter-
mittent regime is clear as the slip ratio increases to unrealistic levels. At a constant value of one,
the slip ratio prediction of the homogeneous model is clearly too low. Lacking any dependence
on quality, the Zivi model appears to be too simplistic. The Baroczy, Premoli, Tandon, and
Yashar models have similar slip ratio trends, generally increasing with quality.
RESULTS AND DISCUSSION
To verify the component models accuracy, predicted heat transfer rates were compared to
measured values in Figure 5. For the condensers, modeled results match within 1.2% of the mea-
sured rate of heat transfer. For the evaporators, most of the modeled results match within 3.3%
of the measured rate of heat transfer. The model underpredicts the measured rate of heat transfer
in the evaporator by as much as 6.1% for the 5 ton (17.6 kW) rooftop unit under HT test
conditions.
Figure 6 shows the measured charge inventory in the 5 ton (17.6 kW) rooftop unit for sub-
cooling. Charge inventory increases nearly linearly with subcooling, and shows only slight
dependence on environmental conditions. This lack of dependence results from mass conserva-
tion. If charge inventory is substantially increased, the added mass must be stored as liquid. In
conventional air conditioners, the refrigerant can only exist as single-phase liquid in the liquid
line and in the condenser. Assuming that subcooling is positive, this added liquid must go into the
1
1.25 Re
L
0.088
X
tt
1 2.85X
tt
0.523
+
---------------------------------------

\
]
| |
2
= Re
L
1125 < X
tt
1 < ,

1
1.2 1.2
1 x
x
-----------
\
| |

g

f
-----
\
| |
0.35
Fr
----------
+ +
---------------------------------------------------------------------
=

K
1
1 x
x
-----------
\
| |

g

f
-----
\
| |
+
-----------------------------------------
=
Harms474.fm Page 69 Monday, December 16, 2002 8:10 AM
70 HVAC&R RESEARCH
Figure 4. Slip Ratio Predictions for Various Void Fraction Correlations
Figure 5. Comparison of Measured and Modeled Rate of Heat Transfer
in Heat Exchangers
Harms474.fm Page 70 Monday, December 16, 2002 8:10 AM
VOLUME 9, NUMBER 1, JANUARY 2003 71
condenser. The associated system responses include increased pressure and heat transfer in the
condenser. As the rate of condensation increases, more of the heat exchanger is dedicated to sub-
cooling the liquid. For the most part, these changes occur irrespective of environmental condi-
tions. The evaporator can hold some of the added refrigerant, but in a system with a thermal
expansion valve, the evaporator superheat is maintained within a tight range and the mass in the
evaporator remains fairly constant. Thus, subcooling is a useful charge inventory parameter for
systems with a thermal expansion valve. The minimum charge can be defined as yielding zero
subcooling. Moderate subcooling can be defined as the subcooling needed to achieve an opti-
mum coefficient of performance. Because increased condenser pressure negatively affects the
compressor power, the optimum level of subcooling occurs over a fairly narrow range, regardless
of the size of the system, for a given refrigerant. With this understanding, a system can be quickly
identified as under- or overcharged.
Total charge inventory predictions for the 2.5 ton (8.8 kW), 5 ton (17.6 kW), and 7.5 ton
(26.4 kW) systems at test condition Aare given in Figures 7, 8, and 9, respectively, as a function
of measured subcooling. In each figure, these predictions are compared to the total measured
charge. All models match the trend of increasing charge with increasing subcooling fairly well,
but there are some distinctions in magnitude. The Hughmark charge inventory results are 5 to
18% higher than the measured values. The Baroczy model slightly underpredicts charge, except
at low subcooling, where the predicted charge is 5% higher than the measured values. The Pre-
moli et al., Zivi, Yashar et al., and Tandon et al. models are clustered together just below the
measured values. At moderate to high subcooling this groups underprediction ranges from 2 to
10%. The homogeneous model yields results 11 to 23% lower than the measured charge.
The relative magnitude of the charge predictions for void fraction models is consistent with
the slip ratio results in Figure 4. The Hughmark methods overprediction is expected because the
slip ratio is overpredicted in the annular regime. Underprediction of the homogeneous model is
Figure 6. Charge as Function of Subcooling for 5 Ton (17.6 kW) Rooftop Unit
Harms474.fm Page 71 Monday, December 16, 2002 8:10 AM
72 HVAC&R RESEARCH
Figure 7. Charge Inventory Results for 2.5 Ton (8.8 kW) Split System at
Test Condition A
Figure 8. Charge Inventory Results for 5 Ton (17.6 kW) Rooftop Unit at
Test Condition A
Harms474.fm Page 72 Monday, December 16, 2002 8:10 AM
VOLUME 9, NUMBER 1, JANUARY 2003 73
also expected because the slip ratio will always be greater than one. The slip ratio results of the
slug flow model suggest that the Premoli et al., Yashar et al., and Tandon et al. models underpre-
dict the slip ratio in the intermittent regime, which leads to a modest underprediction of the
charge inventory. The Zivi models accuracy would suggest that overprediction of the slip ratio
in the intermittent regime is balanced by underprediction of the slip ratio at high quality. Note
that the Baroczy model matches the slip ratio predictions of the mechanistic slug flow model in
the intermittent regime and the mechanistic annular flow model of Tandon et al. at high quality.
The present results suggest that, if a void fraction model works well for one system, then it will
work well with all systems, assuming that the encountered two-phase flow regimes have been
considered and are well represented. Though not presented, charge inventory results at other test
conditions exhibit trends similar to the results at test condition A. This weak dependence on the
environmental conditions is expected from the results shown in Figure 6.
Figures 10, 11, and 12 show the percentage of the total charge inventory by location for the
2.5 ton (8.8 kW), 5 ton (17.6 kW), and 7.5 ton (26.4 kW) systems, respectively. The results are
for test condition A obtained using the Baroczy void fraction correlation. Recall that adding
charge to a system increases subcooling. Although all locations increase in charge as subcooling
increases, most of the added charge goes into the condenser, which has the only percentage of
total charge that increases with subcooling. The increases in charge are nominal in the liquid line
and filter-drier. For locations that are completely liquid, charges increase only because of
increases in density. These are typical results for all the units. Again, the results in Figure 6 sug-
gest a weak dependence on environmental conditions.
Many observations about the relative importance of a given location within the system de-
pend on the arrangement of the system (split system or rooftop unit) and the type of compressor
(reciprocating or scroll). The filter-drier for the 2.5 ton (8.8 kW) split system was a large
Figure 9. Charge Inventory Results for 7.5 Ton (26.4 kW) System at
Test Condition A
Harms474.fm Page 73 Monday, December 16, 2002 8:10 AM
74 HVAC&R RESEARCH
Figure 10. Percentage of Total Charge Inventory by Location for 2.5 Ton (8.8 kW)
Split System at Test Condition A Using Baroczy Correlation
Figure 11. Percentage of Total Charge Inventory by Location for 5 Ton (17.6 kW)
Rooftop Unit at Test Condition A Using Baroczy Correlation
Harms474.fm Page 74 Monday, December 16, 2002 8:10 AM
VOLUME 9, NUMBER 1, JANUARY 2003 75
aftermarket unit and held up to 13% of the total charge (see Figure 10); the one for the 7.5 ton
(26.4 kW) split system was a small OEM unit, holding only 4% of the total charge (see Figure
12). The 5 ton (17.6 kW) rooftop unit did not have a filter-drier. As the liquid and vapor line sets
are added during installation, split systems are more likely than rooftop units to require a filter-
drier. The liquid line location is primarily important in split systems because it must span the
distance between the indoor and outdoor units. Because the 5 ton (17.6 kW) rooftop unit was a
packaged system, the liquid line was rather short.
Vapor in the compressor shell, and refrigerant dissolved in the oil, are important charge loca-
tions to consider in the 7.5 ton (26.4 kW) split system because of the size of the compressor shell
and the oil required to fill it. The 5 ton (17.6 kW) rooftop unit was the only system with a scroll
compressor. For the same size system, the shell of a scroll compressor is much smaller than that
of a reciprocating compressor. As such, vapor in the compressor shell and refrigerant dissolved
in the oil are of secondary importance for the 5 ton (17.6 kW) rooftop unit.
In all systems, the evaporator is important, and its importance is fairly constant. The percent-
age of charge in the evaporator ranged from 12 to 17%. Across all three systems, the condenser
is the most important component to consider. At zero subcooling, the condenser exit is at a satu-
rated liquid state. As subcooling increases, this saturated location moves into the condenser, and
the condenser fills with liquid. Thus, increasing the charge inventory can almost be seen as fill-
ing the condenser.
CONCLUSIONS
The most important issue in charge inventory modeling is proper selection of the void fraction
correlation. Some of the often-used correlations have been shown to be inappropriate based on
flow regime. The most basic charge inventory models use the homogeneous void fraction, but
Figure 12. Percentage of Total Charge Inventory by Location for 7.5 Ton (26.4 kW)
Split System at Test Condition A Using Baroczy Correlation
Harms474.fm Page 75 Monday, December 16, 2002 8:10 AM
76 HVAC&R RESEARCH
this is not justified because some slip between phases will always exist. For the present results,
the homogeneous model underpredicted charge inventory by 11 to 23%. The Hughmark correla-
tion has been used in the past because of its perceived accuracy; however, this model is based on
the bubbly flow regime and, as such, gives an effective slip ratio that is unrealistically high for
air-conditioning applications. In the present study, the Hughmark model overpredicted charge
inventory by 5 to 18%. The Baroczy void fraction correlation gave the best agreement with mea-
sured data. However, the Baroczy correlation is in tabular format and is not especially conve-
nient to apply. If a simple model is required, the Zivi model is recommended: it is as easy to
implement as the homogeneous model, yet predicts system performance much more accurately.
Accurate modeling of heat transfer and pressure drop in the condenser must be achieved. If
the location in the condenser at which the refrigerant becomes saturated liquid is not precisely
determined, then an accurate charge inventory cannot be determined. Besides the heat exchang-
ers, important locations to consider are the line sets, the filter-drier, the compressor shell, and
the refrigerant dissolved in the oil. The liquid line is the most important of these secondary vol-
umes. For split systems, the liquid line is especially important, even more so than the evapora-
tor. When a reciprocating compressor is used, the compressor shell and the refrigerant dissolved
in the oil are important locations to consider. The remaining line sets are less important, but are
not negligible.
NOMENCLATURE
A area, m
2
c
p
specific heat, J/(kg K)
c
p,m
specific heat of moist air, J/(kg K)
c
s
saturation specific heat, J/(kg K)
C
min
minimum heat capacity rate, W/K
C
r
heat capacity ratio
D diameter, m
F friction factor
Fr Froude number
Fr
TD
modified Froude number
Ft Froude rate parameter
G gravitational acceleration, m/s
2
G mass flux, kg/(m
2
s)
H heat transfer coefficient, W/(m
2
K)
h
a,i
inlet air enthalpy, J/kg
h
r,s
saturation air enthalpy at T
ref
, J/kg
h
s,s,o
saturation air enthalpy at tube surface, J/kg
k thermal conductivity, W/(m K)
K flow parameter
L length, m
M mass, kg
mass flow rate, kg/s
m
f
fin parameter
n correlated exponent
NTU number of transfer units
P pressure, Pa
rate of heat transfer, W
Re
L
liquid Reynolds number
S slip ratio
t
c
fin thickness, m
T temperature, C
T
s,o
saturated outlet air temperature, C
u velocity, m/s
UA overall conductance, W/K
V volume, m
3
w
t
fin pitch
We
L
liquid Weber number
X quality
X
tt
Lockhart-Martinelli parameter
Greek
void fraction
effectiveness

f
fin efficiency

o
overall fin efficiency
viscosity, kg/(m s)
refrigerant-oil solubility
density, kg/m
3
surface tension, N/m
two-phase frictional pressure drop multiplier
Subscripts
acc acceleration
air air
cond condenser
evap evaporator
f liquid or fin
fr friction
g vapor
i inner
in inlet
o outer
out outlet
m

Harms474.fm Page 76 Monday, December 16, 2002 8:10 AM


VOLUME 9, NUMBER 1, JANUARY 2003 77
ref refrigerant
s surface
sub condenser subcooling
super evaporator superheat
t total
w wet
REFERENCES
ASHRAE. 1988. Methods of Testing for Rating Unitary Air-Conditioning and Heat Pump Equipment.
ASHRAE Standard 37. Atlanta: American Society of Heating, Refrigerating and Air-Conditioning
Engineers, Inc.
ASHRAE. 1995. Methods of Testing for Rating Seasonal Efficiency of Unitary Air-Conditioners and Heat
Pumps. ASHRAE Standard 116. Atlanta: American Society of Heating, Refrigerating and Air-Condi-
tioning Engineers, Inc.
Bankoff, S.G. 1960. A Variable Density Single-Fluid Model for Two-Phase Flow with Particular Refer-
ence to Steam-Water Flow. Journal of Heat Transfer 82:265-272.
Baroczy, C.J. 1965. Correlation of Liquid Fraction in Two-Phase Flow with Application to Liquid Metals.
Chemical Engineering Progress Symposium Series 61(57):179-191.
Braun, J.E., S.A. Klein, and J.W. Mitchell. 1989. Effectiveness Models for Cooling Towers and Cooling
Coils. ASHRAE Transactions 95(2):164-174.
Carey, V.P. 1992. Liquid-Vapor Phase-Change Phenomena. Washington, D.C.: Hemisphere.
Cavallini, A., D. Del Col, L. Doretti, G.A. Longo, and L. Rossetto. 2000. Heat Transfer and Pressure Drop
During Condensation of Refrigerants Inside Horizontal Enhanced Tubes. International Journal of
Refrigeration 23:4-25.
Cavallini, A. and R. Zecchin. 1974. A Dimensionless Correlation for Heat Transfer in Forced Convection
Condensation. Heat Transfer 1974: Proceedings of the Fifth International Heat Transfer Conference
3:309-331.
Chen, J.C. 1966. Correlation for Boiling Heat Transfer to Saturated Fluids in Convective Flow. Industrial
and Engineering Chemistry Process Design and Development 5:322-329.
Damasceno, G.S., P.A. Domanski, S. Rooke, and V.W. Goldschmidt. 1991. Refrigerant Charge Effects on
Heat Pump Performance. ASHRAE Transactions 97(1):304-310.
Domanski, P. and D. Didion. 1983. Computer Modeling of the Vapor Compression Cycle with Constant
Flow Area Expansion Device. Building Science Series 155. National Bureau of Standards, Washing-
ton, D.C.
Gnielinski, V. 1976. New Equations for Heat and Mass Transfer in Turbulent Pipe and Channel Flow.
International Chemical Engineering 16:359-368.
Haaland, S.E. 1983. Simple and Explicit Formulas for the Friction Factor in Turbulent Pipe Flow. Journal
of Fluids Engineering 105:89-90.
Harms, T.M. 2002. Charge Inventory System Modeling and Validation for Unitary Air Conditioners. Ph.D.
dissertation, Purdue University, West Lafayette, Indiana.
Hughmark, G.A. 1962. Holdup in Gas-Liquid Flow. Chemical Engineering Progress 58:62-65.
Incropera, F.P. and D.P. DeWitt. 1996. Fundamentals of Heat and Mass Transfer, 4th ed. New York:
Wiley.
LeRoy, J.T., E.A. Groll, and J.E. Braun. 2000. Evaluating the Accuracy of PUREZ in Predicting Unitary
Equipment Performance. ASHRAE Transactions 106(1):200-215.
Lockhart, R.W. and R.C. Martinelli. 1949. Proposed Correlation of Data for Isothermal Two-Phase,
Two-Component Flow in Pipes. Chemical Engineering Progress 45:39-48.
Marques, M.E. and C. Melo. 1993. Assessing the Capability of the HPSIM Program to Simulate Room Air
Conditioners. ASHRAE Transactions 99(2):309-316.
Mulroy, W.J. and D. Didion. 1985. Refrigerant Migration in a Split-Unit Air Conditioner. ASHRAE Trans-
actions 91(1A):193-206.
Premoli, A., D. Di Francesco, and A. Prina. 1971. Una Correlazione Adimensionale per la Determinazione
della Densit di Miscele Bifasiche. La Termotecnica 25:17-26.
Rice, C.K. 1987. The Effect of Void Fraction Correlation and Heat Flux Assumption on Refrigerant
Charge Inventory Predictions. ASHRAE Transactions 93(1):341-367.
Harms474.fm Page 77 Monday, December 16, 2002 8:10 AM
78 HVAC&R RESEARCH
Rice, C.K. and W.L. Jackson. 1994. PUREZThe Mark V ORNL Heat Pump Design Model for Chlo-
rine-Free, Pure and Near-Azeotropic Refrigerant Alternatives. Oak Ridge, Tennessee: Oak Ridge
National Laboratory.
Schmidt, T.E. 1949. Heat Transfer Calculations for Extended Surfaces. Refrigerating Engineering
4:351-357.
Taitel, Y. and D. Barnea. 1990. Two-Phase Slug Flow. Advances in Heat Transfer 20:83-132.
Taitel, Y. and A.E. Dukler. 1976. A Model for Predicting Flow Regime Transitions in Horizontal and Near
Horizontal Gas-Liquid Flow. American Institute of Chemical Engineers Journal 22:47-55.
Tandon, T.N., H.K. Varma, and C.P. Gupta. 1985. A Void Fraction Model for Annular Two-Phase Flow.
International Journal of Heat Mass Transfer 28:191-198.
Yashar, D.A., M.J. Wilson, H.R. Kopke, D.M. Graham, J.C. Chato, and T.A. Newell. 2001. An Investiga-
tion of Refrigerant Void Fraction in Horizontal, Microfin Tubes. International Journal of HVAC&R
Research 7(1):67-82.
Zivi, S.M. 1964. Estimation of Steady-State Steam Void-Fraction by Means of the Principle of Minimum
Entropy Production. Journal of Heat Transfer 86:247-252.
Harms474.fm Page 78 Monday, December 16, 2002 8:10 AM

Vous aimerez peut-être aussi