2016 08 03CYCLE - D HXclean PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

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

net/publication/317126407

Refrigerant Performance Evaluation Including Effects of Transport Properties


and Optimized Heat Exchangers

Article  in  International Journal of Refrigeration · May 2017


DOI: 10.1016/j.ijrefrig.2017.05.014

CITATIONS READS

4 123

4 authors:

Riccardo Brignoli J. Steven Brown


The Catholic University of America The Catholic University of America
20 PUBLICATIONS   286 CITATIONS    100 PUBLICATIONS   1,613 CITATIONS   

SEE PROFILE SEE PROFILE

H. M. Skye Piotr A. Domanski


National Institute of Standards and Technology National Institute of Standards and Technology
19 PUBLICATIONS   290 CITATIONS    86 PUBLICATIONS   1,911 CITATIONS   

SEE PROFILE SEE PROFILE

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

Mixed-Gas Joule Thompson Cryoprobe View project

The 23rd Conference on Process Integration, Modelling and Optimisation for Energy Saving and Pollution Reduction - PRES'20 View project

All content following this page was uploaded by H. M. Skye on 21 September 2018.

The user has requested enhancement of the downloaded file.


2016-07-26

Refrigerant Performance Evaluation Including Effects of Transport Properties and


Optimized Heat Exchangers

Riccardo Brignolia, J. Steven Brownb, H. Skyea, Piotr. A. Domanskia*


a
National Institute of Standards and Technology, Gaithersburg, MD 20899, USA
b
The Catholic University of America, Washington, DC 20064, USA
*
Corresponding author, email: [email protected]

Abstract

Preliminary refrigerant screenings typically rely on using cycle simulation models involving
thermodynamic properties alone. This approach has two shortcomings. First, it neglects transport
properties, whose influence on system performance is particularly strong through their impact on the
performance of the heat exchangers. Second, the refrigerant temperatures in the evaporator and condenser
are specified as input, while real-life equipment operates at imposed heat sink and heat source temperatures;
the temperatures in the evaporator and condensers are established based on overall heat transfer resistances
of these heat exchangers and the balance of the system.

The paper discusses a simulation methodology and model that addresses the above shortcomings. This
model simulates the thermodynamic cycle operating at specified heat sink and heat source temperature
profiles, and includes the ability to account for the effects of thermophysical properties and refrigerant mass
flux on refrigerant heat transfer and pressure drop in the air-to-refrigerant evaporator and condenser.
Additionally, the model can optimize the refrigerant mass flux in the heat exchangers to maximize the
Coefficient of Performance. The new model is validated with experimental data and its predictions are
contrasted to those of a model based on thermodynamic properties alone.

1. Introduction

Since the mid-1980s, the refrigeration and air-conditioning sector has devoted a significant effort to identify
and implement refrigerants that address environmental problems of stratospheric ozone depletion and
global warming. The process of selecting the best refrigerants involves consideration of several criteria
(McLinden and Didion, 1987) with the Coefficient of Performance (COP) being of dominant importance
once other ‘gate’ attributes (e.g., chemical stability or lack of toxicity) have been accepted.

Domanski and McLinden (1992) discussed the merits and shortcomings of various methods for predicting
the performance of refrigerants operating in vapor compression cycles. They placed these methods in five
categories from theoretical analysis to laboratory equipment testing: (1) Carnot cycle analysis, (2) simple
methods based on fundamental observations and principles, (3) theoretical and semi-theoretical cycle
analysis, (4) detailed equipment simulation models, and (5) laboratory tests of the vapor compression
equipment. In the refrigerant selection process, great reliance is placed on methods of category (3) for
selecting best candidate fluids for further examination either by more sophisticated models or tests in actual
equipment.

The evaluation methods of category (3) range from idealized thermodynamic cycle analysis to simulations
including some forms of representation of a practical cycle, e.g., non-isentropic compression, refrigerant
pressure drops in the heat exchangers and connecting tubing, or the temperature difference between fluids
exchanging heat. Most often, category (3) methods employ the refrigerant’s thermodynamic properties
alone and perform simulations based on specified evaporator saturation temperature and superheat,
condenser saturation temperature and subcooling, and compressor isentropic efficiency [e.g., CYCLE_D
2016-07-26

model (Brown et al., 2012)]. These models are very popular among refrigeration practitioners because they
are simple and easy to use.

The CYCLE11 model (Domanski and McLinden, 1992) is more advanced conceptually than CYCLE_D-
type models; instead of using evaporator and condenser saturation temperatures specified as input,
CYCLE11 establishes the thermodynamic cycle – including the saturation temperatures – using temperature
profiles of heat-transfer fluids (HTFs) in the evaporator and condenser and respective mean effective
temperature differences between these fluids and the refrigerant, ∆Thx. This approach allows for accounting
for the effect of non-linear temperature glides of zeotropic mixtures during the evaporation and
condensation processes. A simulation example with a high-glide R22/123 mixture showed a COP difference
as high as 8.7 % when the nonlinearity is neglected during simulation.

The cycle simulation can be further advanced by accounting for effects of refrigerant mass flux on the heat
transfer and pressure drop. These effects take place in conventional equipment (air conditioners,
refrigeration systems) employing air-to-refrigerant, fined-tube evaporators and condensers, which rely on
refrigerant forced-convection heat transfer. In these heat exchangers the refrigerant flows through circuits
formed by tubes connected in a serpentine pattern. The designer must choose the number of parallel circuits,
which determines the refrigerant mass flux in these circuits. A higher mass flux improves the refrigerant
heat-transfer coefficient but also increases the refrigerant pressure drop. These two phenomena have
opposite effects on the system COP; the highest COP is achieved with the mass flux that provides the best
balance between the benefit of improved heat transfer and the penalty related to pressure drop.

The effect of refrigerant mass flux on two-phase heat transfer and pressure drop has been noted in the
literature. Domanski and Yashar (2006) compared the performance of five different refrigerants in a vapor
compression system with evaporator and condenser circuitries optimized by a learnable evolution model.
They found that the COP difference between low-pressure and high-pressure fluids changes in favor of
high-pressure fluids when the performance comparison is done using optimized heat exchangers as opposed
to the theoretical cycle evaluation at fixed saturation temperatures in the evaporator and condenser. A
related study showed high-pressure refrigerants to be the best performing fluids in condensers with
optimized refrigerant circuitries (Domanski and Yashar, 2007).

Cavallini et al. (2010) introduced the ‘penalty factor’ and the ‘total temperature penalization’ (TTP)
parameters that combine exergy losses due to heat transfer and pressure drop. They showed that these
parameters can be used either to select the optimum number of circuits in a finned-tube condenser of fixed
overall geometry and number of tubes, or to compare the refrigerants' heat transfer performance in
condensation. Brown et al. (2012) applied this approach to finned-tube evaporators. Further, Brown et al.
(2014) used the ‘penalty factor’ and TTP to compare the evaporation and condensation performance of
several low-GWP refrigerants. Recently, Zilio et al. (2015) compared the seasonal performance of an air-
to-water heat pump working with R410A and R32. They found that a system with an optimized refrigerant
circuitry in the air-to-refrigerant heat exchanger had a 5 % improved seasonal efficiency over a system with
a ‘non-optimal’ circuitry design.

The present paper discusses in detail the effects of refrigerant thermophysical properties and refrigerant
mass flux on refrigerant two-phase heat transfer and pressure drop. Also, the paper discusses a new cycle
model, referred to as CYCLE_D-HX, which accounts for these effects and can select the optimal refrigerant
heat exchanger circuity for maximizing system COP. Examples of CYCLE_D and CYCLE_D-HX
simulations results are presented and contrasted.
2016-07-26

2. Effect of Refrigerant Mass Flux on Two-Phase Heat Transfer and Pressure Drop

A pressure drop in the evaporator or condenser causes a saturation temperature drop and results in an
increased mean effective temperature difference, ∆Thx, between the refrigerant and the external heat transfer
fluid (HTF). The refrigerant heat-transfer coefficient h also affects ∆Thx; a higher h yields lower ∆Thx. A
decrease in condenser and evaporator ∆Thx lowers the temperature lift seen by the compressor and
subsequently lowers the compressor work, which improves the COP.

The functional dependencies of the refrigerant heat-transfer coefficient h and pressure drop ∆p are given in
Eqs. 1 and 2.

, RSF, TPP, , , , (1)

∆ , , RSF, TPP, , , (2)

RSF stands for ‘refrigerant-side feature’ such as an in-tube insert or microfin surface, and TPP stands for
‘thermophysical property’. The heat-transfer coefficient h and pressure drop ∆p are both a function of the
refrigerant mass flux G, where h approximately increases proportionally to G0.8, while ∆p increases
approximately to G2. There is an optimum value of G that best compromises between the two opposite
effects that h and ∆p have on the system COP. This optimum G depends on thermodynamic and transport
properties of the refrigerant, tube diameter, and refrigerant-surface features (RSF in Eqs. 2 and 3).

Figure 1 shows the dependency of h and ∆p on the mass flux in evaporation for R32, a high-pressure fluid,
and R1234yf, a medium-pressure fluid, at Tsat = 0 °C. We used correlations by Wojtan et al. (2005a and
2005b) and Muller-Steinhagen and Heck (1986) for generating data for h and ∆p, respectively. The showed
results are normalized by nominal h and ∆p values for R32 at G=100 kg·m-2·s-1, hnom=3.16 kW·m-2·K-1 and
∆pnom =1.26 kPa, respectively. All values are average values calculated by Eq. 3:

" (3)
! !

where f(x) is h or ∆p, xStart is the quality corresponding to the isenthalpic expansion from a condenser
saturation temperature of 45 °C and 3 K subcooling to evaporator conditions of Tsat=0 °C and xEnd=1.

Figure 1. Normalized forced-convection evaporation heat-transfer coefficient and pressure drop for R32
and R1234yf as a function of mass flux (D=7 mm, smooth tube, Tsat = 0 °C, q=10 kW·m-2)
2016-07-26

Figure 1 shows an important trend of exponentially increasing pressure drop and asymptotically increasing
heat-transfer coefficient with increasing mass flux. R1234yf has a lower heat-transfer coefficient and a
higher pressure drop than R32 at the same mass flux, and the rate of increase of heat-transfer coefficient in
relation to pressure drop is less favorable for R1234yf than for R32. Also, the drop of saturation temperature
of R1234yf is greater than that of R32 for the same pressure drop. For example, at Tsat=0 °C,
dT/dp=0.0934 K·kPa-1 for R1234yf and dT/dp=0.0384 K·kPa-1 for R32, which leads to a 2.4 times greater
decrease in saturation temperature for R1234yf than for R32 at the same pressure drop. Similar trends take
place in condensing flows.

The optimum refrigerant mass flux G is different for each refrigerant and depends on refrigerant
thermodynamic and transport properties. Ammonia, for example, which has a very high thermal
conductivity and moderate saturation temperature drop due to its low vapor density, realizes the best
compromise between heat-transfer coefficient and pressure drop at a low mass flux, i.e., at a high number
of tube circuits. On the other hand, R32 being a higher-pressure fluid (higher vapor density) and having
lower liquid conductivity than ammonia, realizes the best compromise between heat transfer and
temperature drop penalization at a higher mass flux, i.e., at a lower number of circuits.

3. Effect of Thermophysical Properties on Two-Phase Heat Transfer and Pressure Drop

The thermodynamic and transport properties that affect the heat-transfer coefficient, h, and pressure drop,
∆p, in two-phase flow are:
- liquid thermal conductivity, λl
- liquid and vapor density, ρl and ρv
- liquid and vapor viscosity, νl and νv
- liquid specific heat at constant pressure, cp,l
- surface tension, σ (only for evaporation).
The liquid conductivity is the most influential parameter affecting the heat-transfer coefficient, while the
pressure drop is independent of both the liquid and the vapor thermal conductivity. Literature correlations
for h include a term proportional to λla, where a is an exponent in the neighborhood of 0.6; a higher liquid
conductivity results in better two-phase heat transfer. The effect of vapor thermal conductivity is negligible
because of the high flow turbulence; it only slightly modifies the heat-transfer coefficient component for a
dry surface when the evaporation flow is stratified.

The liquid and vapor density also affect the two-phase heat-transfer coefficient: higher ρl yields a larger h,
while higher ρv yields a lower h (for correlations including ρv). The effect of vapor density is greater on the
pressure drop than on the heat-transfer coefficient. The vapor quality strongly affects the two-phase flow
density, and subsequently the pressure drop. Furthermore, a high vapor density corresponds to a small
saturation temperature drop associated with the pressure drop, as shown by the Clapeyron relation
(Borgnakke and Sonntag, 2012):

#$ $&'(
#% )*+
,- − - 1 (4)
. 0

The effect of liquid viscosity on two-phase heat transfer varies depending on the refrigerant and operating
conditions, whereas the effect of vapor viscosity is small. The pressure drop increases as liquid and vapor
viscosities increase. The liquid specific heat affects the heat-transfer coefficient, since it affects the liquid
heat transport capacity. In most correlations h is proportional to cp,la, where ‘a’ is an exponent in the
neighborhood of 0.8. Pressure drop is not a function of cp. The surface tension affects the nucleate boiling
component of the evaporation heat-transfer coefficient; a low surface tension enhances h, but its effect is
small compared with other transport properties.
2016-07-26

In summary, the most important refrigerant properties for the forced-convection heat transfer are the liquid
conductivity and the vapor density: the higher their values, the higher the two-phase heat-transfer
coefficient is. The dominant property affecting pressure drop is the vapor density, where a high value results
in low pressure drop, ∆p, and low saturation temperature drop, dT/dp. Properties of refrigerants differ
significantly (Table 1) so there is a spectrum of refrigerant performance relative to heat transfer and pressure
drop. For example, R32 has both favorable liquid conductivity and vapor density in comparison to most
fluids, whereas ammonia has a low vapor density, which is compensated for by an outstanding liquid
conductivity.

Table 1. Thermophysical properties of common refrigerants at 0 °C dew-point temperature (Lemmon et


al., 2013)

Tc Pc λl ρl νl cp,l ρv νv dT/dp
Fluid
°C MPa W·m-1·K-1 kg·m-3 mPa·s kJ·kg-1·K-1 kg·m-3 mPa·s K·kPa-1
R410A 71.4 4.90 0.103 1170.0 0.161 1.52 30.6 0.012 0.039
R32 78.1 5.78 0.145 1055.3 0.150 1.75 22.1 0.012 0.038
R1234yf 94.7 3.38 0.071 1176.3 0.208 1.29 17.7 0.010 0.093
R134a 101.1 4.06 0.092 1294.8 0.267 1.34 14.4 0.011 0.094
(1)
R717 132.3 11.3 0.559 638.6 0.170 4.62 3.5 0.009 0.062
(2)
R600a 134.7 3.63 0.099 580.6 0.199 2.28 4.3 0.007 0.180
(1)
ammonia; (2) isobutane

4. Description of CYCLE_D-HX Model

4.1 Overall modeling approach


The CYCLE_D-HX1 model builds on the concept of using of temperature profiles of the heat sink and heat
source, and ∆Thx for the evaporator and condenser (Domanski and McLinden, 1992), which facilitates the
accounting for refrigerant thermophysical properties, pressure drop, and heat-transfer coefficient on the
cycle performance on a relative basis (Brown at al., 2002a and 2002b). This section provides a brief
presentation of the CYCLE_D_HX modeling approach with a reference to the source publications. For
simplicity of presentation, we will only discuss the basic vapor compression cycle although the simulation
capabilities of CYCLE_D-HX include enhanced cycle options such as a liquid-line/suction-line heat
exchanger, economizer, and intercooler.

In its simplest form, the simulated system consists of a compressor, condenser, adiabatic expansion device,
and evaporator. The cycle and key thermodynamic states for this system are shown on a temperature-
entropy (T-s) diagram (Figure 2). The compressor is represented by the isentropic efficiency, volumetric
efficiency, and the electric motor efficiency. The evaporator and condenser can be either counter-flow,
cross-flow or parallel-flow, and are represented by their ∆Thx. The solution sequence starts with estimated
values of saturation temperatures in the evaporator and condenser. Based on the established thermodynamic
cycle with refrigerant temperature profiles and HTF temperature profiles, the model calculates ∆Thx and
compares them to the values specified as input. The model iterates evaporator and condenser saturation
temperatures until it achieves the specified ∆Thx values within a convergence parameter.

1
The program will be publically available packaged with a graphical user’s interface and user’s guide.
2016-07-26

Figure 2. Basic vapor compression cycle

For each iteration step of saturation temperatures, CYCLE_D-HX calculates heat exchangers’ ∆Thx using
Eq. 5 (Domanski and McLinden, 1992).
56 59 5<
7 5 … 5 ∑5 (5)
2$34 534 2$6 34 2$9 34 34 2$<

In this equation, ∆Thx is a harmonic mean weighted with the fraction of heat transferred in individual
sections of the heat exchanger, based on the assumption of a constant overall heat-transfer coefficient
throughout the heat exchanger. Each term represents the contribution of a heat exchanger section. At the
outset of each saturation temperature iteration, the model calculates ∆Thx based on sections corresponding
to the subcooled liquid, two-phase, and superheated regions. Then, the model bisects each section and uses
Eq. 5 to calculate a new value of ∆Thx. The model repeatedly bisects each subsection until the ∆Thx obtained
from two consecutive evaluations agree within a convergence parameter.

As an alternative to specifying ∆Thx, the heat exchangers can be characterized by the overall heat
conductance UAhx. If this input option is used, the model calculates the specified ∆Thx from the basic heat-
transfer relation, ∆Thx = Qhx/UAhx, where Qhx is the product of refrigerant mass flow rate and enthalpy
change in the evaporator or condenser, as appropriate.

Representation of heat exchangers by their UAhx allows for inclusion of heat transfer and pressure drop
characteristics in comparable evaluations of different refrigerants. For this purpose, CYCLE_D-HX
considers that the total resistance to heat transfer in a heat exchanger, Rhx, consists of the resistance on the
refrigerant side Rr, and combined resistances of the heat exchanger material and HTF [Rtube + RHTF]:

Rhx = 1/UAhx= Rr + [Rtube + RHTF] (6)

where Rr = 1/( hr·Ahx) (7)

The resistances [Rtube + RHTF] are independent of the refrigerant, and are assumed to be independent of
operating conditions. Their combined value can be calculated from UAhx and hr values using performance
measurements obtained in a laboratory on a system of interest. CYCLE_D-HX calculates [Rtube + RHTF]
within its ‘reference run’ and stores its value for use in subsequent simulation runs for calculation of UAhx
characterizing the heat exchanger with a new refrigerant or operating conditions.
2016-07-26

CYCLE_D-HX requires the following operational input data for the ‘reference run’: HTF inlet and outlet
temperatures for the evaporator and condenser; ∆Thx for the evaporator and condenser (to achieve the
measured evaporator and condenser saturation temperatures); evaporator superheat and pressure drop; and
condenser subcooling and pressure drop. Additional ‘reference run’ inputs include compressor isentropic
and volumetric efficiencies, and electric motor efficiency. Heat exchanger geometry inputs include the
tube inner diameter and length, the number of refrigerant circuits, and the number of tubes per circuit. By
user’s choice, CYCLE_D-HX can optimize evaporator and condenser circuitries (number of parallel
circuits) to maximize the system’s COP. (This represents a design environment where the HTF and number
of refrigerant tubes remains constant, but the refrigerant flow and tube circuitry can be adjusted.) Using this
option, the model provides information on the relative performance potentials of refrigerants operating in
systems with serpentine air-to-refrigerant heat exchangers.

For smooth tubes, CYCLE_D-HX uses correlations by Wojtan et al., (2005a and 2005b) and Shah (2009)
for calculating the forced-convection heat-transfer coefficient for evaporation and condensation,
respectively. For enhanced tubes, the model applies hr correction presented by Shlager et al. (1989). The
modeling of evaporator and condenser refrigerant pressure drop relies on a similar concept to that for the
heat-transfer process. For smooth tubes, the model determines a pressure multiplication factor by dividing
the ‘reference run’ pressure drop by the predicted value (Eq. 8).

∆%!BC
=>?@A∆% ∆%D!B
(8)

where ∆ppred is calculated by the Muller-Steinhangen and Heck (MSH) (1986) correlation for smooth tubes.
For enhanced tubes the MSH value is corrected according to Choi et al. (2001).

Regarding the compressor isentropic efficiency, CYCLE_D-HX offers the option of accounting for its
dependence on the compression ratio, as it was postulated by several researchers (Brown et al., 2002b).
When screening different refrigerants, the model uses Eq. 9 to take into account the change in isentropic
efficiency with the pressure ratio in a consistent way.

E F − 0.05J (9)

Equation (9) has the same slope as the relation derived from experimental data by Brown et al. (2002b); C
is a constant calculated within the ‘reference run’ using the isentropic efficiency and the pressure ratio
obtained from the test data.

4.2 Optimization of refrigerant circuitry


The refrigerant circuitry optimization algorithm identifies the number of condenser and evaporator circuits
that maximize COP. The evaporator is optimized before the condenser because the system is much more
sensitive to the number of evaporator circuits (Zilio et al., 2015); the heat-transfer coefficient and pressure
drop are more sensitive to G in the evaporator because the refrigerant density is lower and the viscosity is
higher. Optimizing the evaporator before the condenser enables more rapid identification of the
multivariate optimum.

The COP optimization algorithm is a modified Golden Section Search, which takes into account the values
and trend. The Golden Section algorithm has been modified because it normally requires a non-flat and
continuous function, but the computed COP function often does not satisfy these criteria. The function may
be flat, for example, when varying the number of tube circuits in the condenser. Discontinuities can occur
from numerical approximations or from the semi-empirical correlations used to evaluate the heat-transfer
coefficient and the pressure drop. For example, Figure 1 shows a discontinuity in the correlation for heat
transfer coefficient in the neighborhood of G=150 kg·m-2·s-1, where the transition from a stratified-wavy to
2016-07-26

an annular flow takes place according to the flow-pattern map of Wojtan et al. (2005a). The algorithm starts
searching in the direction of ascent until an inflection point is found and the maximum resides between the
last two iterations. Then the algorithm makes at least two more iterations and stops when the absolute
difference between the maximum and its two adjacent values is less than the convergence parameter.

The processing times for a simulation with optimization on personal computer (3.5 GHz processor) are on
the order of 10 s, 150 s, and 350 s respectively for one-, two-, and three- component working fluids.

5. CYCLE_D-HX Validation

5.1 Experimental apparatus and test conditions


A laboratory liquid-to-liquid heat pump apparatus was used to generate data to verify the model. The
apparatus (Figure A.1, in Appendix) is described briefly here; a more comprehensive description is
provided in Skye (2015). The system was equipped with a variable-speed reciprocating compressor,
variably-sized evaporator and condenser, manually adjusted throttling valve, and a liquid-line/suction-line
heat exchanger, which could be included or bypassed. The evaporator and condenser were of the annular
design arranged in the counter-current configuration; the refrigerant flowed in the enhanced inner tube
(copper), while the HTF flowed in the smooth annular space. The heat exchangers’ size could be adjusted
by changing the number of active refrigerant tubes; this feature enabled control of the heat flux.

The apparatus was set to achieve evaporation and condensation saturation temperatures nominal to air-
source heat pumps operating at the Cooling A, Cooling B, and Heating H1 rating tests (AHRI, 2008).
Table 2 shows the HTF inlet and outlet temperatures used to obtain these evaporation and condensation
temperatures using R134a and a mid-range compressor speed, 1800 rev·min-1. Four additional data sets at
each rating test (total of twelve) were generated by holding the HTF inlet temperature constant as the system
capacity was varied via compressor speed, (1400 to 2200) rev·min-1; these additional data sets are listed in
the Appendix (Table A1(a), (b), and (c)). Note that the heat exchanger sizes were fixed for the data
presented here.

Table 2. Test operating conditions for R134a tests with 1800 rev·min-1 compressor speed
Capacity Condenser Evaporator
Rating test Tsat,ave THTF,in THTF,out Tsat,ave THTF,in THTF,out
kW
°C °C °C °C °C °C
Cooling A* 1.69 43.6 33.9 39.0 8.1 20.2 15.4
Cooling B 1.95 35.8 24.9 30.9 5.5 19.8 14.2
Heating H1 1.88 39.6 32.0 36.3 0.3 10.2 6.5
*’
reference run’ for CYCLE_D-HX model

Care was taken to configure other evaporator and condenser operating conditions (beyond refrigerant
saturation temperature) to closely resemble those of a typical air-to-air heat pump. Specifically, the heat
fluxes were within (5 to 9) kW·m-2 and (5 to 10) kW·m-2 for the evaporator and the condenser, respectively.
Additionally, the ratios of HTF thermal resistance to total heat exchanger thermal resistance were nominally
0.8 and 0.6 for the evaporator and condenser, respectively; these values are representative of air-to-air heat
pumps where the air-side (i.e. HTF side) thermal resistance dominates. The thermal resistance ratios were
enforced by the selection of HTF mass flow rates; the HTF mass flow rates were held constant for all tests
at 0.098 kg·s-1 for the condenser and 0.131 kg·s-1 for the evaporator. The subcooling and superheat were
controlled to (2 to 3) K and (3 to 6) K, respectively. More details about these tests, including the uncertainty
calculation (95 % confidence level) for the COP (0.35 %), capacity (0.2 %), and the Qvol (1.5 %) are
presented in Skye (2015).
2016-07-26

5.2 Experimental data vs. simulation results


We used the data from the Cooling A R134a test (Table 2) to carry out the CYCLE_D-HX ‘reference run’.
The ‘reference run’ inputs included the 1.69 kW capacity, the evaporator ∆Thx=8.8 K, the condenser
∆Thx = 9.02 K, and pressure drops of 33 kPa and 45 kPa for the condenser and evaporator, respectively. We
then executed simulations of the remaining Cooling A, Cooling B and Heating H1 ratings tests. The
capacities, compressor isentropic and volumetric efficiencies, superheat and subcooling, discharge and
suction line pressure drops, and HTF inlet and outlet temperatures were input based on measurements from
each test.

We evaluated the percentage deviation between the simulation and the experimental results using Eq. 10.
L 4DB!<MB 0 L <MN0 <O
K ⋅ 100 % (10)
L 4DB!<MB 0

where Π is any parameter of interest.

Figure 3 reports the deviations for COP, Qvol, pevap, and pcond. Most of the deviations are within 4 %. The
largest deviation (7.4 %) is for the Cooling B test at the highest (2200 rev·min-1) compressor speed; this
operating condition yielded about 20 % increase in refrigerant mass flow rate and capacity over the
‘reference run’. The model inputs (Table A2(a), (b)) and results (Table A3(a) and (b)) are tabulated in the
Appendix.

Figure 3. Deviations between experimental and CYCLE_D-HX simulations results

6. Comparison of Refrigerant Performance Evaluation Using CYCLE_D and CYCLE_D-HX

For performance evaluation we selected R134a, R600a, and R32 because they have significantly different
thermophysical properties (Table 1). R600a is the lowest-pressure refrigerant of the three, and has a low
vapor density and liquid conductivity. R32 is the highest-pressure refrigerant and has a high liquid
2016-07-26

conductivity and vapor density. The operating pressure, vapor density and liquid conductivity of R134a
are between those of R600a and R32. We simulated the cycle performance of the three fluids at the Cooling
A test operating condition (Section 5.1). The CYCLE_D-HX simulations (without and with the
optimization option) were based on the CYCLE_D-HX ‘reference run’, which applied no optimization and
used the R134a Cooling A test data. We conducted these simulations at the same evaporator capacity that
was specified for the ‘reference run’ (1.69 kW). For all CYCLE_D simulations we used the evaporator and
condenser saturation temperatures and saturation temperature drops from the ‘reference run’.

Figure 4 presents COP results normalized by the ‘reference run’ COP. CYCLE_D predictions correlate
inversely with refrigerants’ Pc: the low-pressure R600a has the highest COP, followed by the medium-
pressure R134a, and the high-pressure R32, which has the lowest COP. The CYCLE_D-HX simulation
results display the opposite trend: the low-pressure R600a shows the lowest COP, and the high-pressure
R32 has the highest COP. The R134a COPs for CYCLE_D-HX without optimization (‘reference run’) and
for CYCLE_D are identical because CYCLE_D simulations used the saturation temperatures and pressure
drop from the ‘reference run’. All three refrigerants benefitted from the refrigerant circuitry optimization.
The COP increased for R600a and R32 by about 5 % and 9 %, respectively. R134a benefitted least, which
suggests that the original refrigerant circuitry was reasonably well designed for R134a.

Figure 4. COP from CYCLE_D and CYCLE_D-HX simulations normalized by COPref = 2.85 (R134a
Cooling A test condition, no optimization)

It is important to recognize the disparity in the COP trends between CYCLE_D and CYCLE_D_HX
simulations, which affects the suitability of these simulation tools for rating competing refrigerants. With
the same (imposed) saturation temperatures in the evaporator and condenser, CYCLE_D simulations tend
to yield higher COPs for low-pressure refrigerants compared to high-pressure refrigerants because low-
pressure fluids operate far below their critical point and tend to exhibit less irreversibilities due to the
superheated vapor horn and throttling process (Domanski and Didion, 1993). However, as shown in the
CYCLE_D-HX simulations, this thermodynamic advantage of low-pressure fluids may be erased in
systems using heat exchangers that implement refrigerant forced-convection evaporation and condensation
2016-07-26

(e.g., air-to-refrigerant coils). Low-pressure fluids have low vapor density and therefore exhibit relatively
large pressure and saturation temperature drops in forced-convection coils. These irreversibilities may
render high-pressure refrigerants more efficient, particularly when refrigerant circuitries (and therefore
mass fluxes) in these heat exchangers are optimized.

The COP trends discussed here are consistent with a prior study, which used a detailed air-to-air heat pump
model with R600a, R134a, R290, R22, R410a, and R32 (Domanski and Yashar, 2006). These performance
trends are also reflected in the engineering practice of selecting low-pressure refrigerants for water chillers
with shell-and tube heat exchangers, which have low pressure drop, and applying higher-pressure
refrigerants in systems with forced-convection heat exchangers. Consequently, we can conclude that the
simulation methodology used by CYCLE_D-HX is preferable for screening refrigerants for systems with
forced-convection heat exchangers over CYCLE_D-type simulation models using imposed saturation
temperatures in the evaporator and condenser, which may provide incorrect ranking of the evaluated fluids.
For systems with insignificant pressure drop, CYCLE_D-type models are likely to produce correct ranking
unless liquid conductivity of the considered fluids differ substantially.

7. Conclusions

Preliminary refrigerant evaluations using theoretical models should consider the type of vapor compression
equipment for which the refrigerant screening is performed. The key aspect differentiating system types is
the type of evaporators and condensers used. For systems with heat exchangers relying on refrigerant pool
boiling, falling film evaporation, or space condensation, cycle models based on thermodynamic properties
alone may provide adequate ‘first estimate’ of relative performance merits of refrigerants of interest.
However, for systems with heat exchangers relying on refrigerant forced-convection evaporation and
condensation heat transfer, more appropriate simulation models are those which also involve refrigerant
transport properties and can account for the effect of refrigerant mass flux on heat transfer and pressure
drop.

REFERENCES

AHRI (2008). 2008 Standard for Performance Rating of Unitary Air-Conditioning & Air-Source Heat
Pump Equipment. Air-Conditioning, Heating, and Refrigeration Institute, Arlington, VA, United States.
Retrieved from www.ahrinet.org.

Borgnakke, C., Sonntag, R. E., 2012. Fundamentals of Thermodynamics, 8th Edition, Wiley, New York.

Brown, J.S., Domanski, P.A, Lemmon, E.W., 2012. CYCLE_D: NIST Vapor Compression Cycle Design
Program, Version 5.0, NIST Standard Reference Database 49, National Institute of Standards and
Technology, Gaithersburg, MD. http://www.nist.gov/srd/nist49.cfm

Brown, J.S., Kim, Y., Domanski, P.A., 2002a. Evaluation of Carbon Dioxide as R22 Substitute for
Residential Conditioning, ASHRAE Transactions, 108(2), 954-963.

Brown, J.S., Yana-Motta, S. F., Domanski, P.A., 2002b. Comparative analysis of an automotive air
conditioning system operating with CO2 and R134a, Int. J. Refrig., 25(1), 19-32.

Brown, J.S., Zilio, C., Brignoli, R., Cavallini, A., 2012. Heat transfer and pressure drop penalization
terms (exergy losses) during flow boiling of refrigerants, Int. J. of Energy Research, 37(13), 1669-1679.
2016-07-26

Brown, J.S., Zilio, C., Brignoli, R., Cavallini, A., 2014. Thermophysical properties and heat transfer and
pressure drop performance potentials of hydrofluoro-olefins, hydrochlorofluoro-olefins, and their blends,
HVAC&R Research, 20, 203-220.
Cavallini, A., Brown, J.S., Del Col, D., Zilio, C., 2010. In-tube condensation performance of refrigerants
considering penalization terms (exergy losses) for heat transfer and pressure drop, Int. J. of Heat and
Mass Transfer, 53, 2885-2896.

Choi, J.Y., Kedzierski, M.A., Domanski, P.A., 2001 Generalized pressure drop correlation for
evaporation and condensation in smooth and micro-fin tube, IIR commission B1 conference,
thermophysical properties and transfer processes of new refrigerants, Paderborn, Germany.

Domanski, P.A., McLinden, M.O., 1992. A Simplified Cycle Simulation Model for the Performance
Rating of Refrigerants and Refrigerant Mixtures, Int. J. Refrig., 15(2), 81-88.

Domanski, P.A., Didion, D.A., 1993. Thermodynamic Evaluation of R-22 Alternative Refrigerants and
Refrigerant Mixtures, ASHRAE Transactions, 99(2), 636-648.

Domanski, P.A., Yashar, D., 2006. Comparable Performance Evaluation of HC and HFC Refrigerants in
an Optimized System, 7th IIR Gustav Lorentzen Conference on Natural Working Fluids, Trondheim,
Norway.

Domanski, P. A., Yashar, D. A., 2007. Optimization of Finned-Tube Condensers Using an Intelligent
System. Int. J. Refrig. 30 (3), 482-488.

Lemmon, E.W., Huber, M.L., McLinden, M.O., 2013. NIST reference fluid thermodynamic and transport
properties-REFPROP, Version 9.1, NIST Standard Reference Database 23, National Institute of
Standards and Technology, Gaithersburg, MD. http://www.nist.gov/srd/nist73.cfm

McLinden, M.O., Didion, A.D., 1987. CFCs: Quest for Alternatives. ASHRAE J., 29(12), 32-42.

Muller-Steinhangen, H., Heck K., 1986. A simple pressure drop correlation for two-phase flow in pipes,
Chem. Eng. and Process., 20, 297-308.

Shah, M.M., 2009.An improved and extended general correlation for heat transfer during condensation in
plain tubes, HVAC&R, 15(5), 889-913.

Shlager, L.M., Pate, M.B., Bergles A. E., 1989. Heat transfer and pressure drop during evaporation and
condensation of R22 in horizontal micro-fin tubes, Int. J. Refrig., 12(1), 6-14.

Skye, H, 2015. Heat Pump Test Apparatus for the Evaluation of Low Global Warming Potential
Refrigerants. Technical Note 1895, National Institute of Standards and Technology, Gaithersburg, MD,
USA. http://dx.doi.org/10.6028/NIST.TN.1895

Wojtan, L., Ursenbacher T., Thome J.R., 2005a. Investigation of flow boiling in horizontal tubes: Part I -
A new diabatic two-phase flow pattern map, Int. J. Heat and Mass Transfer, 48, 2955-2969.

Wojtan, L., Ursenbacher, T., Thome J.R., 2005b. Investigation of flow boiling in horizontal tubes: Part II
- Development of a new heat transfer model for stratified-wavy, dryout and mist flow regimes, Int. J. Heat
and Mass Transfer, 48, 2970-2985.

Zilio, C., Brignoli, R., Kaemmer, N., Bella, B., 2015. Energy efficiency of a reversible refrigeration unit
using R410A or R32, Science and Technology for the Built Environment, 21, 502-514.
2016-07-26
2016-07-26

NOMENCLATURE

A area (m2)
cp specific heat at constant pressure (kJ·kg-1·K-1)
COP coefficient of performance
D diameter (mm)
E deviation (%)
F heat-transfer coefficient or pressure drop
factor∆p pressure drop multiplication factor (dimensionless)
G mass flux (kg·m-2·s-1)
h heat-transfer coefficient (W·m-2·K-1)
hfg enthalpy of vaporization (kJ·kg-1)
HTF heat-transfer fluid
L length (m)
p pressure (kPa)
q heat flux (kW·m-2)
Q cooling or heating capacity (kW)
Qvol volumetric capacity (kJ·m-3)
R heat transfer resistance (K·W -1)
RSF refrigerant side feature (e.g., microfins, insert)
T temperature (K, °C)
TPP thermophysical properties
TTP total temperature penalization (K)
UA heat exchanger conductance (W·K-1)
x vapor quality (decimal fraction)
∆p pressure drop (kPa)
∆Thx mean effective temperature difference (K)
∆T temperature drop (K)

Greek letters
ηs isentropic efficiency (decimal fraction)
θ pressure ratio (dimensionless)
λ thermal conductivity (W·m-1·K-1)
ν kinematic viscosity (mPa·s)
Π any parameter of interest
ρ density (kg·m-3)
σ surface tension (N·m-1)

Subscripts
ave average
c critical
cond condensation
evap evaporation
hx heat exchanger
in inlet
l liquid
nom nominal
2016-07-26

out outlet
prep predicted
r refrigerant
ref reference
sat saturation
tube tube separating refrigerant and HTF
v vapor
2016-07-26

Appendix A: Experimental Data

NOMENCLATURE

cp specific heat at constant pressure (kJ·kg-1·K-1)


COP coefficient of performance
factor∆p pressure drop multiplication factor (dimensionless)
ID inner diameter of tube (m)
LMTD log-mean temperature difference (K)
ST mass flow (kg·s-1)
N speed (rev·min-1)
P pressure (kPa)
QT cooling or heating capacity (kW)
Qvol volumetric capacity (kJ·m-3)
R heat transfer resistance (K·W -1)
RSF refrigerant side feature (e.g., microfins, insert)
SC condenser outlet subcooling (°C)
SH evaporator outlet superheat (°C)
T temperature (K, °C)
∆P pressure drop (kPa)
∆T temperature difference (K)
WT power (kW)
WT UVWX.r compressor power, computed using the refrigerant mass flow and enthalpy change across
compressor (kW)
# number of (tubes, tube circuits, etc.)
1 to 11 thermodynamic states, see Figure A.1

Greek letters
η efficiency (decimal fraction)
τ torque (N·m)

Subscripts
c condenser
comp compressor
cool cooling
controls heat pump controls
dew dew-point (saturated vapor point)
dis discharge line
e evaporator
fan fan for indoor/outdoor heat exchanger
heat heating
HTF heat transfer fluid, or calculated/measured on the HTF side
in inlet
indoor inside the building
ins heat leak to ambient air through HTF outer tube and insulation
m compressor motor efficiency
outdoor outside the building
r refrigerant, or calculated/measured on the refrigerant-side
2016-07-26

s compressor isentropic efficiency


suc suction line
tube tube separating refrigerant and HTF
v compressor volumetric efficiency
1 to 11 thermodynamic states, see Figure A.1
2016-07-26

Figure A1: Schematic of the experimental apparatus


2016-07-26

Table A1(a): Experimental data – refrigerant-side measurements


mT r Ncomp P1 P4 P7 P8 P9 P11 ∆Pc ∆Pe ∆Psuc T1 T3 T4 T7 T8 T9 T11 τcomp
kg·s-1 rev·min-1 kPa kPa kPa kPa kPa kPa kPa kPa kPa °C °C °C °C °C °C °C N·m
* ±0.000016 ±1 ±3.5 ±3.5 ±3.5 ±3.5 ±3.5 ±3.5 ±1.3 ±0.8 ±0.3 ±0.09 ±0.09 ±0.09 ±0.09 ±0.09 ±0.09 ±0.09 ±0.33

0.01037 1426 362.8 1105 1076 1076 412.3 373.7 29.7 38.6 10.9 11.74 72.79 70.04 38.70 37.87 9.76 10.81 4.72
0.01011 1609 370.6 1104 1075 1075 418.4 381.8 29.7 36.7 11.2 14.24 79.15 74.46 38.79 37.91 10.14 13.47 4.08
Cooling 0.01101 1810 353.2 1136 1103 1103 412.0 367.4 33.0 44.6 14.2 12.11 81.92 77.06 38.70 37.83 9.70 11.41 4.25
A
0.01153 2005 337.5 1145 1108 1108 405.8 353.9 37.2 51.9 16.4 12.79 86.52 81.77 39.77 39.12 9.21 12.16 4.27
0.01270 2231 307.8 1170 1127 1127 395.5 329.4 42.3 66.2 21.5 10.47 90.44 87.38 40.99 40.80 8.44 9.66 4.99

0.01094 1429 340 896.9 856.3 856.3 396.3 353.3 40.6 43.1 13.2 9.57 63.49 61.21 31.02 30.73 8.58 8.59 4.72
0.01154 1628 323.8 915.3 869.6 869.6 390.6 340.5 45.8 50.2 16.6 10.51 69.35 66.91 31.56 31.37 8.08 9.64 4.08
Cooling
0.01206 1829 308 932.7 883.9 883.9 384.1 327.1 48.9 57.1 19.1 9.63 73.64 71.12 31.78 31.53 7.62 8.78 4.25
B
0.01246 2032 295.6 946 894.6 894.6 379.8 317 51.4 62.9 21.4 9.79 77.91 75.27 31.91 31.65 7.28 9.02 4.27
0.01291 2234 284.9 958 903.7 903.7 377.6 308.9 54.2 68.7 24.0 8.15 81.12 78.46 32.45 32.27 7.11 7.32 4.99

0.008001 1430 293.1 990.8 968 968 329 300.6 22.7 28.4 7.6 8.02 74.28 70.73 36.38 35.42 3.24 6.46 4.72
0.008433 1628 273.8 1005 979.7 979.7 317.5 284.1 24.9 33.4 10.3 6.80 78.40 74.70 36.66 35.82 2.23 4.98 4.08
Heating
0.009067 1832 264.3 1021 992.8 992.8 315.7 276.5 28.1 39.1 12.3 3.65 81.03 77.40 37.11 36.39 2.05 1.70 4.25
H1
0.009384 2032 253.7 1032 1002 1002 311.2 267.6 29.9 43.6 13.9 4.19 85.69 81.90 37.33 36.63 1.63 2.40 4.27
0.009811 2233 246.9 1046 1014 1014 310.5 262.4 31.9 48.1 15.5 2.28 88.75 84.93 37.45 36.76 1.55 0.37 4.99
*k=2, 95% confidence interval
2016-07-26
Table A1(b): Experimental data – refrigerant-side computed metrics
COPcool COPheat ∆Pdis ηs ηv QT cool.r QT heat.r SC SH Qvol,cool Qvol,heat WT YZ[\.r
-- -- kPa -- -- kW kW °C °C kJ·m-3 kJ·m-3 kW
* ±0.01 ±0.01 ±0.1 ±0.006 ±0.006 ±0.0035 ±0.004 ±0.2 ±0.3 ±30 ±40 ±0.002

3.33 4.26 1.5 0.520 0.562 1.587 2.032 3.39 3.90 2637 3377 0.477
3.07 3.96 1.5 0.468 0.481 1.569 2.027 3.27 5.93 2704 3494 0.512
Cooling 2.82 3.72 1.7 0.463 0.486 1.693 2.235 2567 3388 0.601
4.36 5.00
A
2.62 3.54 1.9 0.457 0.484 1.763 2.378 3.44 6.83 2419 3263 0.673
2.34 3.31 2.3 0.460 0.524 1.889 2.666 3.39 3.90 2152 3038 0.806

3.95 4.92 2.1 0.508 0.629 1.771 2.202 2.69 3.31 2626 3266 0.448
3.57 4.54 2.3 0.497 0.618 1.873 2.38 2.71 5.42 2482 3153 0.524
Cooling
3.24 4.20 2.5 0.483 0.605 1.949 2.532 3.08 5.70 2349 3051 0.602
B
3.02 3.98 2.7 0.475 0.589 2.017 2.663 3.38 6.82 2248 2968 0.669
2.75 3.72 2.8 0.457 0.573 2.061 2.789 3.21 5.86 2147 2906 0.749

3.04 3.97 1.1 0.521 0.538 1.236 1.615 1.79 5.75 2144 2801 0.407
2.76 3.70 1.2 0.511 0.533 1.291 1.732 1.95 5.83 1984 2662 0.468
Heating
2.47 3.43 1.4 0.482 0.521 1.356 1.879 1.99 3.29 1893 2624 0.549
H1
2.33 3.28 1.5 0.473 0.510 1.408 1.985 2.12 4.87 1813 2555 0.605
2.15 3.11 1.6 0.452 0.495 1.454 2.103 2.42 3.38 1755 2537 0.677
*k=2, 95% confidence interval
2016-07-26
Table A1(c): Experimental data – HTF-side measurements Table A1(d): Experimental data – HTF-side computed metrics
mT HTF,c mT HTF,e THTF,c,in ∆THTF,c THTF,e,in ∆THTF,e cp,HTF,c cp,HTF,e QT cool,HTF QT heat,HTF QT ins,HTF,c QT ins,HTF,e
kg·s-1 kg·s-1 K C °C K kJ·kg-1·K-1 kJ·kg-1·K-1 kW kW kW kW
* ±0.00026 ±0.00026 ±0.6 ±0.015 ±0.6 ±0.015 * ±3 % ±3 % ±0.05 ±0.05 ±10 % ±10 %

0.09928 0.1300 33.86 4.625 19.64 4.547 4.18 2.59 1.56 2.002 0.081 0.030
0.09800 0.1317 33.84 4.713 20.20 4.432 4.18 2.59 1.538 2.013 0.081 0.026
Cooling 0.09871 0.1312 33.89 5.158 20.22 4.818 Cooling 1.660 2.213 0.083 0.023
4.18 2.59
A A
0.09801 0.1317 33.88 5.572 20.32 4.976 4.18 2.59 1.724 2.366 0.082 0.027
0.09920 0.1299 33.84 6.092 19.98 5.419 4.18 2.59 1.853 2.613 0.085 0.031

0.09852 0.1300 24.92 5.212 19.72 5.071 4.18 2.59 1.737 2.178 0.030 0.030
0.09861 0.1300 24.82 5.645 19.86 5.356 4.18 2.59 1.838 2.353 0.025 0.036
Cooling Cooling
0.09859 0.1299 24.86 6.007 19.81 5.579 4.18 2.59 1.915 2.502 0.024 0.040
B B
0.09859 0.1299 24.79 6.311 19.90 5.779 4.18 2.59 1.978 2.631 0.029 0.035
0.09846 0.1299 24.88 6.607 19.89 5.922 4.18 2.59 2.025 2.754 0.033 0.034

0.09941 0.1318 32.00 3.643 11.28 3.377 4.18 2.58 1.218 1.584 0.069 0.071
0.09875 0.1317 31.94 3.951 10.37 3.510 4.18 2.58 1.27 1.699 0.067 0.079
Heating Heating
0.09878 0.1317 31.97 4.293 10.23 3.688 4.18 2.58 1.331 1.842 0.069 0.080
H1 H1
0.09872 0.1316 31.91 4.545 10.18 3.847 4.18 2.58 1.385 1.945 0.068 0.082
0.09873 0.1316 31.90 4.831 10.13 3.976 4.18 2.57 1.430 2.063 0.069 0.083
*k=2, 95% confidence interval *k=2, 95% confidence interval
2016-07-26
Table A2(a): CYCLE_D-HX simulations – reference run inputs
Heat Exchangers
Type RSF THTF,c,in THTF,c,out THTF,c,in THTF,c,out LMTDc LMTDe ∆Pc ∆Pe SC SH IDtube Ltube #tube,c #tube,e #cir
-- -- °C °C °C °C K K kPa kPa K K m m -- -- --
Counter Enhanced 33.89 39.04 20.22 15.41 8.46 9.08 32.96 44.64 4.36 5.00 0.00846 0.5588 12 10 1

Compressor Vapor Lines Auxiliary Power


ηs ηv ηm QT cool ∆Tdew,suc ∆Tdew,dis WT fan,indoor WT fan,outdoor WT controls
-- -- -- kW °C °C kW kW kW
0.46 0.49 1.00 1.69 1.14 0.06 0.00 0.00 0.00

Table A2(b): CYCLE_D-HX – reference run results


COPcool Rtube + RHTF factor∆P
-1
-- K·kW --

Condenser 2.794 2.1686 4.329


Evaporator 3.07 4.3507 2.187
2016-07-26

Table A3(a): CYCLED-HX simulations – inputs

Heat Exchangers Compressor Vapor lines Auxiliary Power


Ncomp* THTF,c,in THTF,c,out THTF,e,in THTF,e,out SC SH ηs ηv ηm QT cool ∆Tdew,s ∆Tdew,d WT fan,indoor WT fan,outdoor WT controls
rev·min-1 °C °C °C °C K K -- -- -- kW °C °C kW kW kW

1426 33.86 38.49 19.64 15.09 3.39 3.90 0.520 0.562 1.00 1.587 0.86 0.05 0 0 0
1609 33.84 38.56 20.2 15.77 3.27 5.93 0.468 0.481 1.00 1.569 0.88 0.05 0 0 0
Cooling
A 1810 33.89 39.04 20.22 15.41 4.36 5.00 0.463 0.486 1.00 1.693 1.14 0.06 0 0 0
2005 33.88 39.46 20.32 15.35 3.44 6.83 0.457 0.484 1.00 1.763 1.36 0.06 0 0 0
2231 33.84 39.93 19.98 14.56 3.39 3.90 0.460 0.524 1.00 1.889 1.90 0.08 0 0 0

1429 24.92 30.13 19.72 14.65 2.69 3.31 0.508 0.629 1.00 1.771 1.10 0.08 0 0 0
1628 24.82 30.47 19.86 14.51 2.71 5.42 0.497 0.618 1.00 1.873 1.42 0.09 0 0 0
Cooling
1829 24.86 30.87 19.81 14.23 3.08 5.70 0.483 0.605 1.00 1.949 1.69 0.10 0 0 0
B
2032 24.79 31.11 19.90 14.13 3.38 6.82 0.475 0.589 1.00 2.017 1.95 0.10 0 0 0
2234 24.88 31.48 19.89 13.97 3.21 5.86 0.457 0.573 1.00 2.061 2.24 0.11 0 0 0

1430 32.00 35.64 11.28 7.902 1.79 5.75 0.521 0.538 1.00 1.236 0.70 0.04 0 0 0
1628 31.94 35.89 10.37 6.861 1.95 5.83 0.511 0.533 1.00 1.291 1.01 0.05 0 0 0
Heating
1832 31.97 36.26 10.23 6.546 1.99 3.29 0.482 0.521 1.00 1.356 1.23 0.05 0 0 0
H1
2032 31.91 36.46 10.18 6.336 2.12 4.87 0.473 0.510 1.00 1.408 1.43 0.05 0 0 0
2233 31.90 36.73 10.13 6.158 2.42 3.38 0.452 0.495 1.00 1.454 1.63 0.06 0 0 0
*Compressor speed is not an input to the model, but is listed here to help the reader correlate model inputs to experimental data
2016-07-26
Table A3(b): CYCLED-HX simulations – results
View publication stats

Ncomp* COPcool COPheat Pc,in1 ∆Pc Pe,out2 ∆Pe QT QT cool heat Qvol,cool Qvol,heat WT comp,r
rev·min-1 -- -- kPa kPa kPa kPa kW kW kJ·m-3 kJ·m-3 kW

1426 3.238 4.238 1118 30.64 369.6 40.92 1.59 2.081 2589 3389 0.4911
1609 3.025 4.025 1113 29.29 380.0 37.96 1.57 2.089 2673 3557 0.5189
Cooling
A 1810 2.794 3.794 1136 32.96 367.4 44.65 1.69 2.295 2560 3476 0.6049
2005 2.674 3.674 1146 35.25 360.4 49.24 1.76 2.419 2462 3383 0.6583
2231 2.414 3.414 1168 40.69 338.4 61.07 1.89 2.672 2227 3149 0.7828

1429 3.967 4.967 901.7 41.32 353.5 43.14 1.764 2.208 2634 3298 0.4446
1628 3.678 4.678 911.8 44.4 344.2 48.25 1.867 2.375 2526 3212 0.5077
Cooling
1829 3.339 4.339 925.1 47.29 334.3 53.34 1.947 2.53 2422 3147 0.5831
B
2032 3.151 4.151 932.9 49.49 327.6 57.11 2.017 2.657 2349 3095 0.6402
2234 2.954 3.954 943.7 51.71 322.1 61.24 2.057 2.753 2269 3037 0.6963

1430 2.924 3.924 1016 21.67 298.0 32.93 1.24 1.664 2103 2822 0.4241
1628 2.691 3.691 1024 23.29 282.8 37.36 1.29 1.769 1962 2692 0.4795
Heating
1832 2.409 3.409 1038 26.22 274.1 43.61 1.36 1.925 1869 2645 0.5646
H1
2032 2.295 3.295 1044 27.46 268.2 46.65 1.41 2.024 1814 2605 0.6143
2233 2.135 3.135 1053 29.14 263.1 50.53 1.45 2.129 1760 2585 0.6792
*Compressor speed is not an input to the model, but is listed here to help the reader correlate model inputs to experimental data
1
corresponds to P4 in Figure A.1
2
corresponds to P11 in Figure A.1

You might also like