Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

International Journal of Heat and Mass Transfer 162 (2020) 120301

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/hmt

Condensation of zeotropic mixtures of low-pressure hydrocarbons and


synthetic refrigerants
Srinivas Garimella∗, Jeffrey Milkie, Malcolm Macdonald
Sustainable Thermal Systems Laboratory, George W. Woodruff School of Mechanical Engineering, Georgia Institute of Technology, Love Building, Room 340,
Ferst Drive, Atlanta, GA 30332, United States

a r t i c l e i n f o a b s t r a c t

Article history: Zeotropic mixtures are under consideration to replace pure working fluids for use in Organic Rankine Cy-
Received 4 June 2020 cles for power generation from low-grade heat sources. An experimental investigation of the condensa-
Revised 15 July 2020
tion heat transfer and frictional pressure drop of a zeotropic mixture of R245fa and n-pentane in smooth
Accepted 6 August 2020
horizontal tubes was conducted. These measurements were made over mass fluxes ranging from 150 to
Available online 28 August 2020
600 kg m−2 s−1 , operating pressures from 122 to 610 kPa, and nominal vapor qualities ranging from
Keywords: 0.05 to 0.95. Results from this experimental study are compared with the common zeotropic modeling
Zeotropic mixture condensation approaches using various correlations from the literature.
Experiments
Modeling
© 2020 Published by Elsevier Ltd.
Hydrocarbons
Synthetic refrigerants

1. Introduction a similarly low-pressure fluid with excellent transport properties


that forms a zeotrope in some concentrations with R245fa. An ear-
Geothermal heat sources offer one promising method of power lier study by Milkie et al. [1] investigated the condensation of pure
generation with low carbon footprint. A limiting feature of these R245fa and pure n-Pentane.
cycles is the absence of an ideal working fluid that can efficiently Condensation of a zeotropic mixture is complicated by the non-
utilize the low-grade heat source, and reject to a low-grade heat constant concentrations of each fluid in the two phases. When
sink. Pure fluids can be considered, but there are two significant two fluids with common transport properties are used, the pres-
issues. Firstly, pure fluids suffer from the likelihood of temperature sure drop is largely unaffected by the non-constant concentrations
pinching. Secondly, pure fluids use the temperature difference be- of the two-phases. Therefore, this discussion focuses primarily on
tween the source/sink temperatures and the working fluid temper- the condensation heat transfer. To begin, the heat transfer charac-
ature in an inefficient manner. As such, zeotropic mixtures, which teristics of pure fluids are discussed, followed by a description of
have a non-constant condensation temperature, are considered. As zeotropic mixture condensation heat transfer.
such, zeotropic mixtures, which have a non-constant condensa-
tion temperature, are considered. The non-constant phase change
temperature means the fluid can be tailored to have temperature
1.1. Pure fluid condensation heat transfer
glides to most effectively match the source and sink temperature
changes, thereby minimizing excessive temperature differences and
Pure fluid condensation over a wide range of conditions
the likelihood of temperature pinching. Both points are illustrated
and tube diameters has been studied, resulting in single-regime
in Fig. 1.
[2–4] and multi-regime [5–9] heat transfer correlations. Typically,
R-245fa is one alternative to conventional fluids such as R-134a.
the most accurate correlations attempt to model condensation us-
It has unique thermodynamic and transport properties that make
ing flow regime specific correlations. Thus, these models estab-
it desirable for many industrial applications. Moreover, R-245fa is
lish the dominant heat transfer mechanisms in each flow regime
a “dry” fluid, meaning that expansion across a turbine does not
and propose physically based correlations that predict heat trans-
cause the fluid to condense and form liquid droplets, which are a
fer using an assumed, or known, flow structure. For horizontal
potential source of degradation of the turbine blades. n-Pentane is
tubes, depending on the flow conditions, flow regimes are classi-
fied into intermittent, stratified, annular or mist flow. For condi-

Corresponding author. tions commonly encountered in heavy industry, the most common
E-mail address: sgarimella@gatech.edu (S. Garimella). flow regimes are stratified and annular. As such, most flow regime

https://doi.org/10.1016/j.ijheatmasstransfer.2020.120301
0017-9310/© 2020 Published by Elsevier Ltd.
2 S. Garimella, J. Milkie and M. Macdonald / International Journal of Heat and Mass Transfer 162 (2020) 120301

Condensation in the annular regime is modeled as a circumferen-


Nomenclature tially uniform liquid film.
Jaster and Kosky [11] modeled the stratified regime using a
D tube diameter, m Nusselt film correlation and the heat and momentum analogy of
f friction factor Kosky and Staub [10] in the annular regime. Dobson and Chato
G mass flux, kg m−2 s−1 [6] modeled condensation in the stratified regime using Nusselt
h heat transfer coefficient, W m−2 K−1 film condensation and a forced convection term in the liquid pool.
i enthalpy, kJ kg−1 They proposed a Dittus-Boelter correlation in the annular regime.
k thermal conductivity, W m−1 K−1 Cavallini et al. [7] extended the analysis of Jaster and Kosky [11] to
LMTD log mean temperature difference,°C include forced convection in the liquid pool and proposed using
P pressure, kPa the Kosky and Staub [10] correlation in the annular regime. Thome
Pr Prandtl number, - et al. [9] modeled turbulent flow in the annular regime by mod-
Pred reduced Pressure, - eling the heat transfer in the liquid ring and then proposed an
Q heat duty, W empirical correction factor to account for the increased heat trans-
q vapor quality, - fer due to interfacial roughness. Stratified flow is modeled using a
R thermal resistance, K W−1 combination of Nusselt film theory and the annular regime corre-
T temperature,°C lation, by assuming that the stratified pool is stretched around the
UA heat capacitance rate, W K−1 inner surface of the tube. Cavallini et al. [8] proposed a two-regime
X bulk concentration, kg kg−1 model using empirical adjustments to Nusselt film and liquid pool
x liquid phase concentration correlations in the stratified regime and a single-phase correlation
y vapor phase concentration with a two-phase multiplier in the annular regime.
z distance, m In the study by Milkie et al. [1], the correlations noted above
were compared with condensation heat transfer experiments for
Greek symbols
the pure fluids R245fa and n-Pentane; however, no correlation pre-
α void fraction, -
dicted the measured data well. As such, a new flow regime depen-
 change in variable, -
dent correlation that applied directly to the two pure fluids under
μ viscosity, kg m−1 s−1
investigation in the present study was developed by Milkie et al.
ρ density, kg m−3
[1].
σ surface tension, N m−1

Subscripts/superscripts 1.2. Zeotropic mixture condensation heat transfer


Annulus annulus channel for coolant to flow in test section
Cond conduction through tube walls Zeotropic mixtures exhibit a temperature glide (TGlide = Tdew
fr frictional - Tbubble ) during the phase change process due to differences in
Glide difference bubble and dew point temperatures the boiling points of the two constituent fluids, as demonstrated
L liquid using the T-x-y diagram in Fig. 2. For a mixture of 45% R245fa/55%
meas measured n-pentane (by mass), as shown in Fig. 2, the mixture begins to
sat saturation condense at (B’ –B) at 71.2 °C, and the last vapor is condensed
V vapor at (D’ –D) at 55.0 °C, resulting in a temperature glide of 16.2 °C.
It can also be seen from Fig. 2 that mixtures of two components
can form a zeotropic mixture or an azeotropic mixture, depending
on the concentration of each fluid, but the focus of this study is
dependent correlations are commonly two-regime models that fo- on zeotropic mixtures. When a mixture is zeotropic, mass trans-
cus exclusively on these flow regimes. fer resistances are present, as one fluid preferentially condenses
The dominant thermal resistance to condensation in both these because the two fluids have different boiling points. This added
flow regimes is the liquid thermal resistance. Thus, stratified resistance, which is not present in the case of a pure fluid, has
regime condensation is modeled as a falling film in the upper part the potential to degrade the heat transfer coefficient; therefore,
of the tube, and liquid pool heat transfer in the base of the tube. proper accounting of this effect is required. Fig. 3 illustrates the

Fig. 1. Saturation dome of a zeotropic mixture (left) and pure substance (right), where zeotropic mixture shows non-constant temperature profiles and pinch points.
S. Garimella, J. Milkie and M. Macdonald / International Journal of Heat and Mass Transfer 162 (2020) 120301 3

temperature profiles and heat and mass transfer resistances for


a condensing zeotropic mixture. Fig. 3 demonstrates that as va-
por is condensed, the more volatile component (R-245fa) builds
up at the vapor–liquid interface as the less volatile component
(n-pentane) is preferentially condensed. Thus, in addition to heat
transfer resistances in each flow regime outlined above, a non-
equilibrium condition develops due to the preferential condensa-
tion of the less volatile component. This non-equilibrium condition
causes a mass transfer resistance in the vapor phase that increases
the resistance to phase change. Moreover, preferential condensa-
tion of the less volatile component results in a concentration shift
in both the vapor and liquid phases from the bulk concentration.
This non-equilibrium effect can lead to significant decreases in the
liquid- and vapor-phase temperatures from the expected equilib-
rium temperatures. Furthermore, the temperature difference be-
tween the vapor-liquid interface and the wall are essentially the
driving temperatures for condensation. Thus, lower phase temper-
ature will result in a decrease in the perceived heat transfer rates.
During experimental investigations of the condensation heat trans-
Fig. 2. Equilibrium T-x-y diagram for the 45% R245fa/55% n-Pentane zeotrope. fer coefficient of zeotropic mixtures, these non-equilibrium condi-
tions manifest as an ‘apparent’ reduction in heat transfer coeffi-
cient. A brief overview of the literature pertaining to mixture ef-
fects on the heat transfer process and methods for accounting for
the added resistance using correlations developed for pure fluids is
presented here.
Fronk and Garimella [12] present a review of experimen-
tal studies and modeling techniques for in-tube condensation of
zeotropic mixtures. Most studies focus on refrigerant mixtures for
HVAC applications [13, 14]. Many researchers have observed and
documented apparent heat transfer coefficients of zeotropic mix-
tures to be lower than the value resulting from a linear interpo-
lation of the two pure component values. There are two common
methods in the literature used to account for these mixture effects
on the heat transfer coefficient. The first approach was developed
by Colburn and Drew [15] that provides a framework for model-
ing the respective mass, species and energy transport in the both
the vapor and liquid phases. This method based on the underlying
phenomena correctly models the non-equilibrium conditions that
develop due to the preferential condensation of one of the com-
ponents of a binary mixture. Price and Bell [16] later outlined a
method that could be used to implement this approach in the de-
sign of heat exchangers. The second approach was developed by
Silver [17] and later generalized by Bell and Ghaly [18]. This ap-
proach is essentially an equilibrium approach that approximates
the mass transfer resistance that develops in the vapor phase by
over-estimating the heat transfer resistance in the vapor phase. The
resulting thermal resistance is then used with a pure fluid conden-
sation heat transfer correlation in series to estimate the condensa-
tion behavior of the zeotropic mixture. Typically, the heat transfer
coefficient predicted using this method is called the ‘apparent’ heat
transfer coefficient.
Experimental studies of hydrocarbon/hydrocarbon, hydrocar-
bon/water, and refrigerant/refrigerant zeotropic mixtures have
compared the predictions of these two methods with the mea-
sured heat transfer coefficients. The fluid mixture studied has been
found to significantly affect the agreement of correlations from
the literature with experimental results. Price and Bell [16] found
that for condensation of mixtures of 43% methanol/57% water
and 18% n-butane/82% n-octane in a 25.4 mm tube, the approxi-
mate method proposed by Bell and Ghaly [18] tends to underesti-
mate the measured heat transfer coefficient and over predicts the
concentration of the more volatile component in the vapor bulk.
Similarly, they found that when used in heat exchanger calcula-
Fig. 3. Concentration and temperature profiles in vapor and liquid phases of con- tions, this method over predicted the heat transfer area required
densing zeotropic mixture with heat and mass transfer resistance networks. for condensation. Hashimoto et al. [19] studied methanol/ethanol,
ethanol/water and ethanol/benzene mixtures in 25 mm downward
4 S. Garimella, J. Milkie and M. Macdonald / International Journal of Heat and Mass Transfer 162 (2020) 120301

flow. They found good agreement for methanol/ethanol mixtures section. The test section is a tube-in-tube heat exchanger with
using pure fluid models for the liquid Nusselt number using the cooling water in counter-flow to the working fluids.
Colburn and Drew [15] method. However, they found poor agree- Heat gains and losses in the primary loop, and the rest of the
ment in the zeotropic portion of the methanol/water mixtures us- facility, are estimated using a one-dimensional resistance network.
ing the same approach, attributing the poor agreement to inaccu- To minimize ambient losses, low thermal conductivity insulation
rate knowledge of the fluid mixture mass transfer properties. Cav- (0.04 W m−1 K−1 ) was used to wrap the entire experimental fa-
allini et al. [7] found good agreement between their data on a con- cility, including pure fluid, mixture, and coolant lines. These calcu-
densing binary mixture of R-125 and R-236ea and the heat duty lated heat losses are incorporated into all heat duty calculations
predicted using the Price and Bell [16] method using the Cavallini and are assigned a conservative 50% uncertainty. To make sure
and Zecchin [20] correlation. that heat losses are not a major factor, an energy balance is con-
Del Col et al. [21] proposed a slight modification to the Silver ducted on the facility by summing all heat rejected to the coupling
[17], Bell and Ghaly [18] method to account for interface and non- fluid lines and comparing it with the heat added to the loop from
equilibrium effects. Cavallini et al. [8] included zeotropic mixture the evaporator. Excellent agreement was achieved between these
data in the development of their pure fluid heat transfer model. quantities for all data reported in this manuscript.
They indicate that their correlation can be used within the Price Bulk mixture concentration measurements are made using an
and Bell [16] framework, but recommend using the more easily im- in-line gas chromatograph (GC) (Shimadzu Scientific GC-2014,
plemented approximate method proposed by Bell and Ghaly [18] to ±0.0023 concentration). Located between the post-condenser out-
correct their pure fluid correlation. Wen et al. [22] measured heat let (state point (4) in Fig. 5) and the pump, the GC is equipped
transfer coefficients for propane, butane, and propane/butane mix- with a thermal conductivity detector (TCD) using helium as the
tures condensing in small diameter tubes. They found good agree- carrier gas. A small portion of the subcooled liquid is extracted
ment with the correlation of Dobson and Chato [6] with 100% of from the working fluid loop through an automated 100 μL liquid
the data being predicted within ±20%. sampling loop. The sample is transported from the valve to sample
Condensation studies in the literature have focused primar- loop using helium and carried to the GC injector. The sample is
ily on pure refrigerants, pure hydrocarbons, azeotropic, and near- passed throughout the GC column (ResTek Rt-Alumina BOND/CFC)
azeotropic mixtures. Mixtures of dissimilar fluids, such as hydro- where the two components are separated. The separated sample
carbons and refrigerants, are absent from the literature; therefore, then passes through the TCD detector, resulting in two distinct
a comprehensive investigation of binary fluid mixture condensa- peaks measuring the concentration of the liquid sample. These two
tion of 45% R245fa/55% n-Pentane (by mass) was conducted in the peaks are transformed to concentrations using the Shimadzu Lab-
present study at typical power plant condenser conditions. This Solutions V5.42 SP2 [26] software package.
specific combination of fluids exhibits ~16 °C temperature glide The thermal amplification technique is used to measure the
between the dew and bubble points over the range of conditions heat transfer coefficient with low uncertainties. The primary
of interest, which facilitates glide matching between the work- coolant loop that is connected to the test section operates at a high
ing fluid and the external heat source/sink fluids. Therefore, in flow rate, which results in a low thermal resistance, Rannulus , in the
the present study, condensation heat transfer coefficients and fric- UA-LMTD calculation, Eq. (1).
tional pressure gradients for this mixture inside a circular 7.75-
Q˙ test = UA · LMT D
mm smooth tube were measured over the entire condensing re-  −1 (1)
gion (nominally 0.05 < q < 0.95), over a range of mass fluxes = 1
hcondensation A
+ RCond + RAnnulus · TLM
(150, 300, 450 and 600 kg m−2 s−1 ) and saturation temperatures
High flow rates typically lead to small changes in the coolant
(TBubble/Sat = 30 °C, 55 °C and 80 °C).
temperature, which results in high uncertainties in the measured
heat transfer coefficient. Thus, a secondary loop is coupled to the
2. Experimental approach
high flow rate primary coolant loop that operates at a very low
flow rate. The low flow rate causes a large temperature rise in
The test facility used for the experiments in the present study
the coolant in the secondary loop. The combination of the accu-
is shown in Fig. 4. This facility was originally used for flow visual-
rate measurement of the heat duty using the secondary loop and
ization and pressure drop experiments by Coleman and Garimella
the low thermal resistance of the coolant to the UA-LMTD calcula-
[23] and Garimella et al. [24], and subsequent heat transfer exper-
tion in the primary loop minimizes uncertainties in the heat trans-
iments by Bandhauer et al. [25] for the condensation of R-134a in
fer coefficients. Equilibrium temperatures at the inlet and outlet of
channels with hydraulic diameters in the range 0.4 to 5 mm. In the
the test section are used in the LMTD calculation, Eq. (2); thus, the
present work, additional modifications were made to this facility
heat transfer coefficients that are presented in the present study
to facilitate zeotropic mixture experiments, such as the addition of
are apparent heat transfer coefficients.
temperature probes throughout the facility to measure liquid- and
   
vapor-phase temperatures. Detailed descriptions of the test facil- TEq.−in − Tcoolant−out − TEq.−out − Tcoolant−in
ity design, system operation and data reduction methods are pre- TLM =   (2)
ln
(TEq.−in −Tcoolant−out )
sented in the papers above. (TEq.−out −Tcoolant−in )
Fig. 4 shows a schematic of the test loop and coupling fluid
loops used for these experiments. A subcooled liquid is pumped A differential pressure transducer measures the total pressure
to a steam-coupled tube-in-tube heat exchanger, where the fluid is drop across the test section. The acceleration pressure drop (decel-
evaporated and superheated. The superheated state is verified us- eration pressure gain) is calculated using Eq.n (3) and used to iso-
ing temperature, pressure and bulk concentration measurements. late the frictional pressure drop contribution. Here, the void frac-
The fluid is partially condensed in the pre-condenser to a desired tion is determined by the correlation developed by Baroczy [27]. A
test section inlet quality. A quantity of heat is removed in the test conservative uncertainty of 25% on the acceleration pressure drop
section, which is measured using the thermal amplification tech- (deceleration pressure gain) is used.
nique outlined in Bandhauer et al. [25]. After leaving the test sec-
Pf r = Pmeas + G2
tion, the fluid is subcooled in the post-condenser. The test section  2   2 
outlet quality is determined using an energy balance between the q (1 − q )2 q (1 − q )2
+ − + (3)
subcooled state and the heat duty removed in the post-condenser ρV α ρL (1 − α ) in ρV α ρL (1 − α ) out
S. Garimella, J. Milkie and M. Macdonald / International Journal of Heat and Mass Transfer 162 (2020) 120301 5

Fig. 4. Test facility schematic with measurement devices shown.

Fig. 5. Gas chromatograph valves and loop used to determine composition of zeotropic mixture.

Table 1 the present study was 11% with a minimum uncertainty of 7% and
Summary of test facility measuring devices and their respective
maximum uncertainty of 18%. The results shown in Fig. 6 are for
measurement uncertainties.
a mass flux of 300 kg m−2 s−1 . The zeotropic mixture trends with
Measuring Device Measurement uncertainty mass flux, quality, and pressure are similar to those of the pure
T-Type Thermocouple ± 0.25 °C fluids. However, as expected, a reduction in heat transfer coeffi-
Absolute Pressure Transducer ± 7 kPa cients of the zeotropic mixture, relative to the pure components,
Differential Pressure Transducer ± 0.088 kPa was observed due to the additional mass transfer resistances. None
Working Fluid Flowmeter ± 0.1%
of the conditions shown in Fig. 6 are identical; however, at approx-
Pre/Post Condenser Flowmeter ± 0.1%
Primary Coolant Flowmeter ± 0.5% imately similar reduced pressures (Pred = 0.04 for n-Pentane and
Secondary Coolant Flowmeter ± 0.15% Pred = 0.03 for R245fa) the n-Pentane has a noticeably higher heat
transfer coefficient at all quality points. This is becauase R245fa has
a lower latent heat and lower liquid-phase thermal conductivity
The data from these experiments are analyzed using Engineer- and higher liquid viscosity at similar reduced pressures. This trend
ing Equation Solver (EES) [28], with fluid properties obtained from persists for higher reduced pressures.
REFPROP Version 9.0 [29]. An uncertainty propagation analysis was Table 2 shows a comparison of the relevant thermodynamic and
conducted for each data point using the Taylor and Kuyatt [30] ap- transport properties for the two pure fluids and the zeotropic mix-
proach in EES. A summary of the uncertainties in the measure- ture. Fig. 7 shows the heat transfer coefficients of the pure flu-
ments used in this study is shown in Table 1. ids and zeotropic mixtures and a comparison with a composition
weighted average from the pure fluids, which is what might be ex-
3. Results and discussion pected if equilibrium conditions existed. Here, the pure fluids were
condensing at 45 °C and compared against the zeotropic mixture
Fig. 6 compares a sample of the experimental results from the with Tbub = 30 °C and Tbub = 55 °C, where the mass flux and vapor
present study with the pure fluid study conducted by Milkie et al. quality are maintained at 300 kg m−2 s−1 and 0.35, respectively.
[1]. The average uncertainty in the heat transfer coefficient from In the absence of the mass transfer resistances, the heat transfer
6 S. Garimella, J. Milkie and M. Macdonald / International Journal of Heat and Mass Transfer 162 (2020) 120301

Fig. 6. Heat transfer coefficient versus quality (from left to right) of pure n-Pentane, pure R245fa, and 45% R245fa/55% n-Pentane mixture.

Table 2
Comparison of relevant thermodynamic and transport properties of pure fluids and zeotropic mixtures.

Pred ρL ρV ρL ρV −1 μL kL σ ifg
Fluid - kg m−3 kg m−3 - kg m−1 s−1 W m−1 K−1 N m−1 kJ kg−1

R245fa 0.03 1352 7 189 4.29 × 10−4 0.090 0.014 193


0.08 1282 17 78 3.08 × 10−4 0.082 0.011 178
0.17 1204 34 36 2.25 × 10−4 0.074 0.008 160
n-Pentane 0.04 601 4 153 1.85 × 10−4 0.104 0.013 350
0.06 585 6 97 1.64 × 10−4 0.099 0.012 337
0.10 568 9 63 1.46 × 10−4 0.094 0.010 323
45% R245fa/ 55% n-Pentane 0.05 794 7 109 2.17 × 10−4 0.098 0.014 285
0.11 754 15 51 1.71 × 10−4 0.089 0.011 267
0.21 709 27 26 1.36 × 10−4 0.080 0.008 244

ity. This is because a larger proportion of the fluid is low-density


vapor, which will increase the mean fluid velocity, thus increasing
interfacial shear and the frictional pressure gradient. The frictional
pressure gradient is highest for the lowest reduced pressure, be-
cause the vapor density is lowest at the lower reduced pressure,
thus directly increasing the frictional pressure gradient, which can
be seen in Eq. (4). Additionally, Table 2 shows that the equilib-
rium thermodynamic properties of the zeotropic mixture are be-
tween those of the two pure fluids. Similarly, the frictional pres-
sure gradient for the mixture at a given set of conditions is be-
tween the two pure fluids. Overall, the frictional pressure gradient
is predominantly influenced by the velocity, Eq. (4), and the vapor
thermodynamic properties, which do not change significantly with
the concentration changes that occur in the test section. Thus, the
frictional pressure gradient does not exhibit the trends observed
for the heat transfer coefficients.
dP 1 v2
= f ρ (4)
dz 2 D
Fig. 7. Representative mass transfer resistance degradation of heat transfer coeffi-
cient.
4. Comparison with the literature

The measured heat transfer coefficients of the 45% R245fa/55%


coefficients of the mixture would be expected to be substantially n-pentane zeotropic mixture are compared with correlations in
closer to the concentration weighted average of the two pure flu- the literature with the correction factor of Bell and Ghaly [18].
ids based on the governing fluid properties and flow conditions. The Bell and Ghaly [18] comparison is followed by a comparison
However, the mixture results show that the mixture heat transfer of the zeotropic mixture measured test section heat duties with
coefficients were significantly closer to the R245fa results, confirm- those predicted using the non-equilibrium Price and Bell [16] ap-
ing the degradation of the heat transfer coefficient due to mass proach. The liquid film heat transfer coefficient is modeled using
transfer resistances. pure fluid condensation models as inputs. Because the Price and
Fig. 8 compares a sample of the frictional pressure gradient re- Bell [16] approach calculates the mass transfer effects specific to
sults, at a mass flux of 300 kg m−2 s−1 , from the present study each set of inlet conditions, operating conditions, and condensation
with the pure fluid study conducted by Milkie et al. [1]. As ex- heat transfer rate, the film heat transfer coefficient can significantly
pected, the frictional pressure gradient increases with vapor qual- alter the calculated outlet conditions. The comparisons between
S. Garimella, J. Milkie and M. Macdonald / International Journal of Heat and Mass Transfer 162 (2020) 120301 7

Fig. 8. Frictional pressure gradient versus quality (from left to right) of pure n-Pentane, pure R245fa, and 45% R245fa/55% n-Pentane mixture.

Fig. 9. Heat transfer coefficient predictions of Milkie et al. [1] model compared with zeotropic mixture measurements using the method by Bell and Ghaly [18] (left) and
the method by Price and Bell [16].

the zeotropic mixture data and correlations from the literature us- The 45% R245fa/55% n-pentane zeotropic mixture at P = 198
ing the Price and Bell [16] approach are best conducted based on kPa (TBub = 30 °C) has a vapor-phase thermal diffusivity of
comparisons of the heat duty. Therefore, the mixture results are 1.50 × 10−6 m2 s−1 and a binary diffusion coefficient in the
compared with the approximate method [17, 18] using the appar- vapor phase, calculated using the Chapmann-Enskog relation, of
ent heat transfer coefficient and the nonequilibrium method [15, 1.5 × 10−6 m2 s−1 . This results in a vapor Lewis number of 0.99. At
16] using the heat duty. P = 763 kPa (TBub = 80 °C) the vapor Lewis number is 0.87. Webb
8 S. Garimella, J. Milkie and M. Macdonald / International Journal of Heat and Mass Transfer 162 (2020) 120301

Fig. 10. Heat transfer coefficient predictions of Cavallini et al. [8] model compared with zeotropic mixture measurements using the method by Bell and Ghaly [18] (left) and
the method by Price and Bell [16] (right).

Fig. 11. Heat transfer coefficient predictions of Thome et al. [9] model compared with zeotropic mixture measurements using the method by Bell and Ghaly [18] (left) and
the method by Price and Bell [16] (right).

et al. [31] found the Bell and Ghaly [18] approach to yield accu- The overall agreement using the correlation presented in Milkie
rate results for fluids with vapor Lewis numbers of 0.6 to 0.8. For et al. [1] is shown in Fig. 9. The comparisons with the meth-
Lewis numbers greater than unity, they suggest that the method ods used to account for condensation of the zeotropic mixture are
will over predict the heat transfer coefficient. Therefore, in this shown at the top of the figure and the equilibrium model predic-
case, the findings of Webb et al. [31] indicate that a slight over tions at experimental conditions are overlaid on the data at the
prediction may result from the use of the Bell and Ghaly [18] ap- bottom of Fig. 9. For the zeotropic mixture, both methods of ac-
proach. counting for the mass transfer resistance contribution to the pre-
The average deviation (AD) and average absolute deviation dicted heat transfer coefficient show good agreement; there is an
(AAD), as defined in Eq. (4), are used to compare the measured AAD of 14% and 8% for the Bell and Ghaly [18] method and Price
and predicted values. and Bell [16] method, respectively. The Bell and Ghaly [18] method
predicts 93% of the data within ±25%, while the Price and Bell
1 [16] method predicts 96% of the data within ±25%. The underlying
AD=
n pure fluid correlation was developed specifically for the fluid con-


1
predicted−measured ditions under investigation in the present study, which results in
n n
predicted−measured
&AAD= the good agreement. Moreover, the Bell and Ghaly predictions do
measured n measured
i =1 i =1 not result in significant over-predictions of the heat transfer coef-
(5) ficient, as predicted by Webb et al. [31].
S. Garimella, J. Milkie and M. Macdonald / International Journal of Heat and Mass Transfer 162 (2020) 120301 9

Table 3
Comparison between measured heat duty and predicted heat duty using correlations from the literature.

Correlation Bell and Ghaly (1973) Price and Bell (1974)


AD, % AAD, % % within ±25% AD, % AAD, %

Thome et al. [9] 27 28 51 28 30


Cavallini et al. [8] 9 15 82 20 22
Psat , kPa
Milkie et al. [1] 198 -13 14 93 -1 7
412 -11 14 0 7
763 -6 15 1 9

Comparisons similar to the Milkie et al. [1] comparison above, CRediT authorship contribution statement
shown in Fig. 9, are conducted using two commonly used pure
fluid correlations by Cavallini et al. [8] and Thome et al. [9], and Srinivas Garimella: Conceptualization, Methodology, Writing -
the results are plotted in Figs. 10 and 11. Figs. 10 and 11 demon- original draft, Project administration. Jeffrey Milkie: Methodology,
strate that using either of these correlations would result in an Validation, Formal analysis, Writing - original draft. Malcolm Mac-
over-prediction of the zeotropic mixture heat transfer coefficient. donald: Formal analysis, Writing - original draft.
This result should not be unexpected; Milkie et al. [1] demon-
strated that these correlations tended to similarly over-predict the References
underlying pure fluid heat transfer coefficients. However, this re-
sult demonstrates the importance of the underlying pure fluid cor- [1] J.A. Milkie, S. Garimella, M.P. Macdonald, Condensation heat
transfer and pressure drop of low-pressure hydrocarbons and
relations that are used in either the Bell and Ghaly [18] or Price synthetic refrigerants, Int. J. Heat Mass Transf. 161 (2020)
and Bell [16] methods for an accurate prediction of the zeotropic https://doi.org/10.1016/j.ijheatmasstransfer.2020.120295.
mixture condensation. [2] D.P. Traviss, W.M. Rohsenow, A.B. Baron, Forced-convection condensation in-
side tubes: a heat transfer equation for condenser design, ASHRAE Trans. 79
Using the Cavallini et al. [8] correlation with the Bell and Ghaly (Part 1) (1973) 157–165.
[18] correction predicted 82% of the data within ±25%, with an [3] M.M. Shah, A general correlation for heat transfer during film condensation
AAD of 15%. Using the Price and Bell [16] approach with the Cav- inside pipes, Int. J. Heat Mass Transf. 22 (1979) 547–556.
[4] H.M. Soliman, The mist-annular transition during condensation and its in-
allini et al. [8] correlation to predict the test section heat duty, the fluence on the heat transfer mechanism, Int. J. Multiph. Flow 12 (2) (1986)
AD and AAD were 20 and 22%, respectively. Using the Thome et al. 277–288.
[9] correlation with the Bell and Ghaly [18] correction, the cor- [5] T.N. Tandon, H.K. Varma, C.P. Gupta, Heat transfer during forced convection
condensation inside horizontal tube, Int. J. Refrig. 18 (3) (1995) 210–214.
relation predicted 51% of the data from the present study within
[6] M.K. Dobson, J.C. Chato, Condensation in smooth horizontal tubes, J. Heat
±25% with an AAD of 28%. Using the Price and Bell [16] approach Transf. 120 (1) (1998) 193–213.
with the Thome et al. [9] correlation to predict the test section [7] A. Cavallini, G. Censi, D.D. Col, L. Doretti, G.A. Longo, L. Rossetto, Condensa-
tion of halogenated refrigerants inside smooth tubes, HVAC R Res. 8 (4) (2002)
heat duty, the AD and AAD were 35 and 36%, respectively. The
429–451.
agreement between all three correlations is summarized in Table 3. [8] A. Cavallini, D.D. Col, L. Doretti, M. Matkovic, L. Rossetto, C. Zilio, G. Censi,
Moreover, because both the Bell and Ghaly [18] and Price and Condensation in horizontal smooth tubes: a new heat transfer model for heat
Bell [16] methods are over-predicted by approximately the same exchanger design, Heat Transf. Eng. 27 (8) (2006) 31–38.
[9] J.R. Thome, J. El Hajal, A. Cavallini, Condensation in horizontal tubes, part 2:
amount for both pure fluid correlations, this result may suggest new heat transfer model based on flow regimes, Int. J. Heat Mass Transf. 46
that the Lewis number approach outlined by Webb et al. [31] may (18) (2003) 3365–3387.
not be applicable to these conditions. [10] P.G. Kosky, F.W. Staub, Local condensing heat transfer coefficients in the annu-
lar flow regime, AIChE J. 17 (5) (1971) 1037–1043.
[11] H. Jaster, P.G. Kosky, Condensation heat transfer in a mixed flow regime, Int. J.
5. Conclusions Heat Mass Transf. 19 (1976) 95–99.
[12] B.M. Fronk, S. Garimella, In-tube condensation of zeotropic fluid mixtures: a
Heat transfer coefficients and frictional pressure gradients were review, Int. J. Refrig. 36 (2) (2013) 534–561.
[13] D.W. Shao, E. Granryd, Experimental and theoretical study on flow condensa-
measured for a zeotropic mixture of 45% R-245fa/ 55% n-Pentane tion with non-azeotropic refrigerant mixtures of R32/R134a, Int. J. Refrig. 21
over a range of conditions applicable to the geothermal power (3) (1998) 230–246.
generation industries in 7.75-mm diameter smooth tubes. Heat [14] S. Koyama, A. Miyara, H. Takamatsu, T. Fujii, Condensation heat transfer of bi-
nary refrigerant mixtures of R22 and R114 inside a horizontal tube with inter-
transfer coefficients, in general, increase with increasing mass flux, nal spiral grooves, Int. J. Refrig. 13 (4) (1990) 256–263.
quality, and decreasing saturation temperature. Pressure gradients [15] A.P. Colburn, T.B. Drew, The condensation of mixed vapours, AIChE Trans. 33
also generally increase with increasing mass flux, quality, and de- (1937) 197–212.
[16] B.C. Price, K.J. Bell, Design of binary vapor condensers using the colburn-drew
creasing saturation temperature. The heat transfer coefficients de- equations, AIChE Symp. Ser. – Heat Transf. – Res. Des. 70 (138) (1974) 163–171.
creased more than would be predicted using a weighted aver- [17] L. Silver, Gas cooling with aqueous condensation, Ind. Chem. Chem. Manuf. 23
age the individual pure fluids. The Bell and Ghaly [18] and Price (269) (1947) 380–386.
[18] K.J. Bell, M.A. Ghaly, An approximate generalized design method for multicom-
and Bell [16] methods were used to predict the behavior of the
ponent/partial condensers, AIChE Symp. Ser. 69 (131) (1973) 72–79.
zeotropic mixture. Both methods perform well when the Milkie [19] R. Hashimoto, K. Yanagi, T. Fujii, Condensation of organic binary mixtures flow-
et al. [1] correlation is used for the underlying pure fluid model, ing downward inside a vertical tube, Heat Transf. – Jpn. Res. 25 (6) (1996)
362–381.
which was developed specifically for these fluids at these condi-
[20] A. Cavallini, R. Zecchin, A dimensionless correlation for heat transfer in forced
tions. The predictions of the zeotropic mixture heat transfer coef- convective condensation, in: Proceedings of the 5th International Heat transfer
ficient are heavily dependent on the underlying pure fluid model. Conference, JSME, 1974, pp. 309–313.
Commonly used correlations from the literature, e.g., Cavallini et al. [21] D. Del Col, A. Cavallini, J.R. Thome, Condensation of zeotropic mixtures in hor-
izontal tubes: new simplified heat transfer model based on flow regimes, J.
[8] and Thome et al. [9], performed poorly for the underlying pure Heat Transf. 127 (3) (2005) 221–230.
fluids and also for the zeotropic mixture heat transfer coefficients. [22] M.-Y. Wen, C.-Y. Ho, Condensation heat-transfer and pressure drop characteris-
tics of refrigerant R-290/R-600a-oil mixtures in serpentine small-diameter U–
Declaration of Competing Interest tubes, Appl. Therm. Eng. 29 (11-12) (2009) 2460–2467.
[23] J.W. Coleman, S. Garimella, Visualization of two-phase refrigerant flow during
phase change, in: Proceedings of the 34th National Heat Transfer Conference,
None. 20 0 0.
10 S. Garimella, J. Milkie and M. Macdonald / International Journal of Heat and Mass Transfer 162 (2020) 120301

[24] S. Garimella, J.D. Killion, J.W. Coleman, Experimentally validated model for [29] E.W. Lemmon, M.L. Huber, M.O. McLinden, Reference Fluid Thermodynamic
two-phase pressure drop in the intermittent flow regime for circular mi- and Transport Properties – REFPROP Version 9.1, National Institute of Stan-
crochannels, J. Fluids Eng. Trans. ASME 124 (1) (2002) 205–214. dards and Technology, 2013 NIST Standard Reference Database 23Standard Ref-
[25] T.M. Bandhauer, A. Agarwal, S. Garimella, Measurement and modeling of con- erence Data Program.
densation heat transfer coefficients in circular microchannels, J. Heat Transf. [30] B.N. Taylor, C.E. Kuyatt, Guidelines for Evaluating and Expressing the Uncer-
128 (10) (2006) 1050–1059. tainty of NIST Measurement Results, National Institute of Standards and Tech-
[26] Shimadzu, LabSolutions, in, 2011. nology, 1994 (SP 1297).
[27] C. Baroczy, Correlation of Liquid Fraction in Two-Phase Flow with Application [31] D. Webb, M. Fahrner, R. Schwaab, The relationship between the colburn and
to Liquid Metals, Atomics International, 1963. silver methods of condenser design, Int. J. Heat Mass Transf. 39 (15) (1996)
[28] S.A. Klein, Engineering equation solver (EES) for Microsoft Windows Operat- 3147–3156.
ing System: academic professional version, in: F-Chart Software, Madison, WI,
2012.

You might also like