Air-Cooled Solar Absorption Air Conditioning
Air-Cooled Solar Absorption Air Conditioning
FINAL REPORT
Engineering Thermodynamics
Mekelweg 2
NL - 2628 CD Delft
Tel. ++31.15.2784894
Telefax ++31.15.2782460
email [email protected]
Report K-SOLAR
Table of Contents
VOORWOORD ........................................................................................................................................................ 7
SUMMARY............................................................................................................................................................... 9
SAMENVATTING................................................................................................................................................. 11
1 INTRODUCTION AND GOALS ................................................................................................................ 13
1.1 BACKGROUND OF THE REPORT ............................................................................................................... 13
1.2 OBJECTIVES ........................................................................................................................................... 13
1.3 OVERVIEW OF THE REPORT .................................................................................................................... 13
2 STATE-OF-THE-ART / POSITION HALF EFFECT CYCLE............................................................... 15
2.1 STATE OF THE ART ................................................................................................................................. 15
2.1.1 Absorption chillers in the market ..................................................................................................... 15
2.1.2 Air-cooled absorption cycles ............................................................................................................ 16
2.1.3 NH3/H2O solar absorption chiller ..................................................................................................... 21
2.2 CONCLUSIONS ........................................................................................................................................ 26
3 DESIGN AND CONSTRUCTION OF HALF-EFFECT LIBR/H2O CHILLER .................................... 29
3.1 THERMODYNAMIC AND TRANSPORT PROPERTIES ................................................................................... 29
3.1.1 A Gibbs energy expression for lithium bromide aqueous solutions ................................................. 29
3.1.2 Transport properties.......................................................................................................................... 31
3.2 THERMODYNAMIC DESIGN OF HALF-EFFECT CHILLER ............................................................................ 35
3.2.1 Introduction ...................................................................................................................................... 35
3.2.2 Working principle of SPHE cycle .................................................................................................... 35
3.2.3 Modeling .......................................................................................................................................... 38
3.2.4 Parametric study ............................................................................................................................... 45
3.3 DESIGN OF CHILLER SETUP ..................................................................................................................... 49
3.3.1 Cycle design ..................................................................................................................................... 49
3.3.2 Chiller configuration......................................................................................................................... 53
3.3.3 Generator .......................................................................................................................................... 57
3.3.4 Condenser ......................................................................................................................................... 59
3.3.5 Low pressure absorber / mid-pressure evaporator ............................................................................ 59
3.3.6 Mid pressure absorber ...................................................................................................................... 60
3.3.7 Low pressure evaporator .................................................................................................................. 61
3.3.8 Single phase heat exchangers ........................................................................................................... 61
3.3.9 Chiller setup...................................................................................................................................... 62
4 FALLING FILM EXPERIMENTS............................................................................................................. 65
4.1 MOTIVATION OF THE STUDY................................................................................................................... 65
4.2 DESIGN OF FILM FLOW SETUP ................................................................................................................. 67
4.3 HEAT AND MASS TRANSFER IN FALLING FILM FLOW ............................................................................... 68
4.4 HEAT TRANSFER COEFFICIENT ON COOLANT SIDE .................................................................................. 70
4.5 FALLING FILM TESTS AND ANALYSIS ...................................................................................................... 75
4.6 CONCLUSIONS ........................................................................................................................................ 82
5 TESTING OF CHILLER SETUP ............................................................................................................... 83
5.1 PRELIMINARY TESTS .............................................................................................................................. 83
5.1.1 Vacuum tightness ............................................................................................................................. 84
5.1.2 Flows from the generator to the absorbers........................................................................................ 85
5.1.3 Recirculation of refrigerant............................................................................................................... 86
5.1.4 Overflowing solution from MPA...................................................................................................... 88
5.1.5 Escaping solution from LPA ............................................................................................................ 89
5.2 TEST RESULTS – 1ST CHARGING CONDITION ............................................................................................ 90
Voorwoord
Zongedreven airconditioning met lucht gekoelde absorptiekoelsystemen
Airconditioning van gebouwen is een snel groeiende energie gebruikende sector. Deze groei is
het gevolg van een toename van behoefte aan comfort en verhoogde levenstandaard in
combinatie met grotere thermische belastingen ten gevolge van kantoorapparatuur. Zonne-
energie kan (tenminste theoretisch) een groot deel van de energiebehoefte voor
airconditioningdoeleinden dekken. Absorptiekoelmachines worden door thermische energie
aangedreven en zijn geschikt voor de inzet in combinatie met thermische zonne-
energiewinning. Probleem is dat relatief hoge aandrijftemperaturen nodig zijn en dat
koeltorens moeten worden ingezet om acceptabele prestaties te bewerkstelligen. Het eerste
probleem vraagt om de toepassing van dure zonnecollectoren terwijl het tweede probleem
extra investeringen vergt.
Doelstelling van dit project is een bijdrage leveren aan de ontwikkeling van een “single-lift
half effect” LiBr/H2O absorptiekoelmachine. In een dergelijke absorptiekoelmachine wordt
een tweede absorber bij een tussendruk ingezet. De lage druk absorber wordt regeneratief
gekoeld. Het concept maakt het mogelijk om relatief lage aandrijftemperaturen in combinatie
met relatief hoge warmteafvoertemperaturen in te zetten. Zo kunnen goedkoop vlakke plaat
zonnecollectoren worden toegepast met behoud van een acceptabel rendement en kan
goedkoop droge luchtkoeling worden ingezet.
Summary
This final report is a summary of R&D activities carried out in the project, ‘Solar Absorption
Cooling’ from March 2003 to May 2005.
Solar-based air conditioning systems allow for significant reduction of the primary energy use
for cooling of buildings. In this report, the efforts developed to come to an economic
competitive concept are discussed.
Since H2O/LiBr is preferred for air-cooled solar absorption systems, its thermodynamic and
transport properties are investigated.
Thermodynamic simulations are used to compare parallel and series versions of the half effect
cycle resulting in specifications for a 10 kW machine.
Development of designs and fabrication methods for construction of a complete chiller are
discussed. Some radical changes have been made in the original designs mainly for financial
reasons. Shell-and-plate generators turned out to be the most expensive component and the
idea of using them was abandoned because the particular type of design is not vital in realizing
the chiller. The condenser design had to be changed because it was part of the generator in the
original design. After considering several alternatives, it was decided to reuse the shell-and-
tube generator used in the NH3/H2O setup together with a new shell-and-coil condenser. The
other parts of the chiller are mostly the same as the original designs except the fact that they
are without extended heat transfer surfaces. These changes bring down the system capacity
substantially but they were inevitable.
Since no extended heat transfer surface was to be implemented in the chiller setup, a small-
scale experimental setup has been constructed to investigate heat and mass transfer processes
in falling films over those surfaces. It includes a small plate absorber inside a glass tube and
the other auxiliary devices to support the absorption process on the absorber surface. The
experimental results obtained with the setup completed the results from the chiller setup.
A chiller model has been developed. Mass transfer resistances in falling film flows have been
included and all equations were expressed in “linear” forms adequate for matrix solution.
Experiments with the half-effect system have been performed under several operating
conditions. The maximum cooling capacity was 4.36 kW, 44% of the original design. The
maximum obtained COP was 0.25, 66% of the design value. The reduced COP mainly
resulted from the design modifications required in relation to costs. However some of this
reduction is caused by problems in fabrication. It should however be remarked that the
operating conditions were quite different from the design conditions. This also has a
significant impact on performance. For operating conditions similar to the design conditions
the capacities and efficiencies would be much smaller than designed.
The development of a model for economic analysis has been initiated. The analysis requires
accurate prediction of the dynamic performance of a system for extended periods. Since the
nature of this task is comparative and therefore simulation has to be carried out for various
systems of different configurations, the analysis program has been structured in such a way
that the system can be easily reconfigured. Attention has also been paid to the model
development in order to reduce the computing time. The half-effect system is assessed in
combination with different solar collector types.
Samenvatting
Dit rapport geeft een overzicht van de werkzaamheden verricht in het kader van het project
‘Zongedreven airconditioning met lucht gekoelde absorptiekoelsystemen’ gedurende de
jaren 2003 tot en met 2005. Enkele onderwerpen die besproken zijn in de voorgaande
voortgangsrapportages worden kort besproken. De niet eerder gerapporteerde onderwerpen
worden uitgebreider behandeld.
Zonnegedreven airconditioning kunnen voor een belangrijk deel bijdragen aan de reductie van
het primaire energiegebruik benodigd voor het koelen van gebouwen. Dit rapport behandelt de
inspanningen die in dit kader zijn verricht om te komen tot een economisch concurrerend
concept.
De parallel en serie geschakelde versies van de “half-effect” cyclus zijn vergeleken op basis
van specificaties voor een 10 kW machine.
Ontwerp en fabricage methoden voor de constructie van een complete “chiller” worden
besproken. Ten opzichte het originele ontwerp zijn een aantal ingrijpende modificaties
toegepast om de kosten te beperken. Toepassing van een “shell-and-plate” generator blijkt tot
onacceptabele kosten te leiden. Ook de condensor moest worden aangepast omdat het een
onderdeel vormde van de generator in het oorspronkelijke ontwerp. Uiteindelijk is een “shell-
and-tube” generator ingezet. De overige warmte- en stofwisselaars zijn uitgevoerd als
spiraalvormige warmtewisselaars met gladde buizen. Deze wijzigingen leiden tot een
significante reductie van de prestaties ten opzichte van het oorspronkelijke ontwerp, maar
waren onvermijdelijk.
Omdat er geen verbeterde oppervlakten zijn toegepast in de chiller setup, is een kleinschalige
experimentele opstelling gebouwd om warmte- en stofoverdracht processen in vallende films
te onderzoeken. Deze opstelling bestaat uit een kleine absorber in een glazen pijp en enkele
apparaten om het absorptieproces in de absorber te ondersteunen. De experimentele resultaten
van deze opstelling dienen als aanvulling op de resultaten verkregen met de chiller
opstellingen.
Experimenten met het half-effect systeem zijn uitgevoerd onder verschillende procescondities.
The maximale koelcapaciteit was 4.36 kW, 44% ten opzichte van het originele ontwerp. De
maximaal haalbare COP was 0.25, 66% van het originele ontwerp. De lagere COP is
hoofdzakelijk veroorzaakt door de wijzigingen in het ontwerp die nodig waren om de kosten
te reduceren. Echter, een deel van de verlaging is ook toe te schrijven aan constructieve
problemen. Daarnaast moet worden opgemerkt dat de procescondities in sterke mate afweken
van de ontwerpcondities. Ook dit heeft een aanzienlijke invloed op de prestaties. Voor
procescondities gelijk aan de ontwerpcondities zouden capaciteiten en rendementen veel lager
dan ontwerpen zijn.
De ontwikkeling van een model ten behoeve van economisch analyse is geïnitieerd. De
analyse vraagt om nauwkeurige voorspelling van de tijdsafhankelijke dynamisch prestaties.
Omdat het model bedoeld is om systemen met verschillende combinaties te vergelijken, is het
programma gestructureerd om aanpassingen mogelijk te maken. Het half-effect systeem is
geëvalueerd in combinatie met verschillende typen zonnecollectoren.
Most of the air-cooled absorption chillers developed in the past, required high driving
temperatures around 120°C. Since flat solar collectors are generally not efficient in this
temperature range, they were equipped with vacuum tube or concentric type collectors.
Consequently, the capital investments were so high that the benefits of primary energy saving
from solar-driven operation did not seem so attractive.
Water-cooled systems, on the other hand, have problems associated with the need of a cooling
tower. The large initial and operation cost is burdensome for small- and medium-sized
applications, let alone the other problems including hygiene and water consumption.
1.2 Objectives
The goal of this project is to develop an air-cooled system with a minimal capital investment
using a low-temperature absorption cycle that enables a chiller to run with economic flat solar
collectors. A target is to develop an air-cooled absorption chiller with 10 kW cooling capacity
that is capable of running at a driving temperature of 90°C in typical mid-summer weather
conditions in southern Europe.
As a result of the thermodynamic design of the proposed system, a complete chiller setup has
been tested. And also another setup, which has been designed for investigation of falling film
flows, has been tested.
Dynamic system models have been developed and some of them are intended especially for
the use in economic analysis of solar absorption cooling systems.
This report gives an overview of the activities developed in the frame of this project.
Chapter 2 presents the state-of-the-art of solar absorption cooling and the position of the half-
effect cycles in this context.
Chapter 3 describes the design and construction of the half-effect cycle setup. It starts by
introducing the thermodynamic and transport properties of LiBr-H2O. A thermodynamic
simulation is then presented on basis of which the cycle components have been designed. The
different components are finally introduced.
Chapter 4 introduces the falling film setup and the results are discussed.
Chapter 5 explains the experimental results obtained with the half-effect cycle setup.
Chapter 8 gives an assessment of the half-effect cycle for solar assisted air conditioning
systems.
Chapter 9 summarizes conclusions from the R&D activities and gives recommendations for
the future.
Until now, no absorption chiller is available in the market for air-cooled solar cooling
application. Table 2.1 summarizes information on small absorption chillers in or near to the
market known to the authors.
Except for the last two gas-fired chillers, all the others require reasonably low heating
temperatures (80~90 oC) that could be attained from a mid-temperature solar collector. But it
should be noted that they are all based on SE (Single-Effect) water-cooled cycles.
Reasons for this domination of SE water-cooled machines in the market are clear. A single
effect machine is simple and easy for fabrication and thus cheap. But since its required heating
temperature is too high for air-cooled operation(>120 oC), it needs expensive high-temperature
solar collectors, which in turn requires the whole solar collector circuit (collectors, storage,
pumps, valves, piping and etc.) to comply with pressure vessel codes if water is to be used as
heating medium. Therefore a cooling tower must be an inevitable option for SE machines.
The gas-fired chillers such as Rinnai and Robur are based on high-temperature cycles like
DE(Double-Effect) and GAX(Generator-Absorber-eXchange), which require heating medium
temperatures above 150 oC for air-cooled operation. The heating temperature requirement of
these high-efficiency cycles is so high as to effectively limit their application even to water-
cooled solar cooling.
Disadvantages of a water-cooled system become clear considering the related initial (cooling
tower, pump, piping and etc.) and the operation cost (water and electricity consumption,
maintenance, sanitation and etc.).
According to the recent survey of European solar cooling [Final report, SACE 2003], average
water consumption of water-cooled solar cooling systems in Europe is 5.3kg/kWh cooling.
Although the price of water may be negligible, this is nevertheless a huge waste considering
that only 1.4 kg of water, as a refrigerant, has an equivalent cooling power (latent heat) of
1kWh. The same source also revealed that average electricity consumption is 225W//kW
cooling, most of which is used for circulation of cooling water.
Although water-cooled system may be a practical solution for high-cooling load applications
such as multistory office buildings, hotels and so on, it is clearly not appropriate for less
demanding applications like residential or small- and medium-size office air conditioning.
This project was motivated by understanding that absence of a small air-cooled machine, such
as targeted in this project, in the market is one of the reasons that keep solar cooling
technologies from claiming its due shares in the air conditioning market.
The content of this section is based on two papers by Kim and Machielsen [2002a, 2002b]. A
purpose of the two studies was to select the best cycle for an air-cooled solar absorption
chiller. Only important results are presented here. More details can be found in the original
papers.
Every absorption cycle has its unique characteristics in regard with its driving (i.e. heating)
temperature and the efficiency. The following is the results of thermodynamic analysis on
various absorption cycles that are appropriate for air-cooled operation.
Among various absorption cycles, the simplest is the so-called SE (Single-Effect) cycle. It has
a moderate generator temperature and efficiency. Its schematic process and P-T-x cycle
diagrams are given in Figure 2.1.
SE cycle has only a few, rather simple recuperative heat exchangers and yields only
marginally acceptable COP for most applications.
Since high-temperature energy sources are often used to drive absorption chillers, some
modifications can be made to the SE cycle to maximize recovery of waste heat for higher
COP. AHX (Absorber Heat eXchange) and GAX (Generator Absorber eXchange) cycles are
typical ones. Both cycles are based on the simple idea that more heat can be recollected when
the temperature of waste heat source is higher. Due to the addition of extra heat exchangers,
cycle is more complicated than SE.
For the energy sources with extremely low temperatures, so-called HE (Half-Effect) cycles
can be used at the expense of reduced COP. Since COP of this kind is about half of the SE
cycle, it is often called half-effect. The generator temperature is lower than SE cycle roughly
by 30 K. There are two basic cycles in this category. DPHE (Double-Pump Half-Effect) and
SPHE (Single-Pump Half-Effect) are introduced below.
DPHE has two separate solution circuits and thus two solution pumps. Refrigerant vapor is
exchanged between mid-pressure generator and absorber. Roughly the same amount of heat is
supplied to each generator to get a cooling effect. SPHE, on the other hand, has only one
solution circuit and a single solution pump. Part of the refrigerant is used to cool down an
intermediate absorber and the rest of the refrigerant is used for cooling effect. But both HE
cycles are similar in performance.
All those cycles above have so-called ‘cut-off’ temperatures, a characteristic generator
temperature below which operation is impossible. To overcome this obstacle, a regenerative
cycle called ‘1R’ cycle was invented by Dao [1990]. This cycle was devised to operate
continuously in all temperature ranges using many absorbers and generators. Although COP is
the highest of all, it is applicable only to large systems because of its complexity.
Simulations were carried out for thermodynamic comparison of those cycles. Below are only
the main results are presented.
1
Carnot 1R
GAX
0.8
0.6 AHX
COP
SE
0.4
DPHE
0.2
0
60 80 100 120 140 160
Tgen ( C) o
COP of each cycle can be seen in Figure 2.6. The required heating medium temperature for
each cycle can also be estimated from the figure. An overall cooling efficiency, η can be
estimated by multiplying the efficiency of a solar collector at the corresponding heating
medium temperature.
Three models of solar collectors were considered to investigate how collector efficiency
changes these results. The first one is a vacuum tube collector that yields high efficiency over
a wide range of temperature. The second is a high-performance flat solar collector with
moderate heating medium temperature in design condition. The last one is a normal flat solar
collector, of which the heating medium temperature is the lowest of all. COP curves in Figure
2.6 were combined with the efficiency curves of the solar collectors and the results are shown
in Figure 2.7.
As it is clear from the figure, AHX and GAX chillers would need vacuum tube collectors to
compete with SE chiller. Considering the potential heat losses and a long warm-up time for
those high-temperature chillers, it was concluded that a target system should be based on a
low-temperature cycle.
40 Vacuum 40
30 30
20 20
10 10
0 0
40 Flat I 40
30 30
η (%) 20 20
10 10
0 0
40 Flat II 40
30 30
20 20
10 10
0 0
60 80 100 120 140 160
Tgen (oC)
Figure 2.7 η vs. Tgen (NH3/H2O, Tamb=35°C, I=800 W/m2)
20
40
water-cooled SE (H2O/LiBr)
water-cooled SE (H2O/LiBr) 16
30
air-cooled
HE (H2O/LiBr)
12
air-cooled HE (H2O/LiBr)
η avg(%)
η avg(%)
20
0 0
4 5 6 7
4 5 6 7
Qsol(kWh/m2day) Qsol(kWh/m2day)
(a) Flat collector I (b) Flat collector II
Figure 2.8 - ηavg vs. Tgen (Tamb=35°C, I=800 W/m2)
As it is shown in Figure 2.8, HE H2O/LiBr chiller is better than SE NH3/H2O chiller for all
cases. And the performance difference is bigger with a low-efficiency collector like Flat II in
Figure 2.8b. Another interesting result is that HE H2O/LiBr can be even comparable to water-
cooled SE H2O/LiBr system if Flat collector II has to be used. The reason for all this results
originates from the low generator temperature and relatively high COP of the HE H2O/LiBr
chiller.
Based on those results above, a half-effect H2O/LiBr cycle was thought the most adequate for
an air-cooled solar absorption cooling system.
This section presents the results of a NH3/H2O program carried out at TU Delft during the
academic years 2002 to 2003. Main objective was to develop an NH3/H2O absorption chiller
by modifying minimum number of the components in an existing gas-fired NH3/H2O heat
pump.
Fernández-Seara and Vásquez [2001] investigated the impact of generator and heat rejection
temperatures for a single stage NH3/H2O absorption refrigeration machine. Since the
evaporating temperature was maintained constant and was relatively low (-15°C) no specific
conclusions can be drawn for solar air conditioning applications from their results. Keizer
[1982] investigated the performance of a single stage NH3/H2O absorption refrigeration
machine and derived a regression line for prediction of performance in the investigated
operating range. Figure 2.9 shows the results obtained when end generation, end absorption
and condensation temperatures are varied while the evaporating temperature is maintained at
0°C. The efficiencies of the different heat exchangers are assumed constant.
0.700
T_condensation/end absorption=30 C
0.600
0.500
0.300
0.000
80 90 100 110 120 130
End generator temperature [C]
These results suggest that operation with heat rejection from absorber and condenser directly
to the surrounding air in most cases is feasible while low-cost flat-plate solar collectors can
operate in a range where they are reasonably efficient. With these results in mind, the idea of
modifying an existing gas-fired NH3/H2O absorption refrigeration machine to a solar energy
driven machine was developed. This paper reports about the modifications carried out,
experimental results obtained with the modified setup and perspectives of the system based on
experiments and model predictions.
Experimental setup
Figure 2.10 Schematic of Figure 2.11 Process diagram of experimental NH3/H2O setup
the generator design
A water-heated generator was designed and the original gas-fired generator was replaced. The
new generator is a shell and tube type heat exchanger where hot water flows across the
vertical tube bank on the shell-side. The strong solution flows as a film inside the vertical
tubes and is heated to generate refrigerant vapor. A distribution system is positioned at the top
of the generator to evenly distribute the solution to the tubes. The schematic design of the
generator is given in Figure 2.10.
The solution circuit was originally an absorber heat exchange (AHX) cycle, Kim and
Machielsen [2002a]. Since this idea is not applicable for low-temperature cycles as required
by solar-based absorption, the solution circuit was modified to form a single effect cycle. The
expansion devices were replaced with metering valves to allow control of weak solution and
refrigerant flows. A process diagram and a photograph of the actual system are given in
Figures 2.11 and 2.12, respectively.
Figure 2.12 Experimental NH3/H2O setup with the generator mounted in the left-side frame.
The setup was charged with 17 kg ammonia-water solution with a weight concentration of
40%. Pressure, temperature and flow instrumentation was installed on the setup. Temperature,
pressure and flow were logged at the inlet and outlet from both sides of each heat (and mass)
exchanger.
Experimental results
Sixty-eight experimental runs have been executed: the first thirty-four with a charging
ammonia water concentration of 40%, the second group with a charging ammonia
concentration of 42.5%.
Table 2.2 indicates the operating range investigated. The values printed bold correspond to the
reference conditions that are maintained as each of the different parameters is varied in the
range shown in Table 2.2.
0.700
charging concentration = 42.5%
0.600
0.400
COP [-]
0.300
0.200
0.100
0.000
20 25 30 35 40
Cooling water flow [lpm]
Figure 2.13 COP of the system as function of the cooling water mass flow rate
At first, the performance of the system has been studied varying the cooling water mass flow
rate through the condenser and absorber for both ammonia concentrations. Figure 2.13 shows
the impact of the cooling water flow rate on the COP. As expected, larger mass flows give
0.600
0.400
COP [-]
0.300
charging concentration = 40.0%
0.200
0.100
0.000
30 32 34 36 38
Cooling water inlet temperature [C]
Figure 2.14 COP of the system as function of the cooling water inlet temperature
The impact of the inlet temperature of the heat rejection medium (water in this case) was
investigated in the range 30 to 38°C. Figure 2.14 shows the impact of the cooling water inlet
temperature on the COP. As expected, lower cooling water inlet temperatures give better
results. Operation up to 38°C is possible with a heating medium temperature of 110°C while
the COP of the system is still acceptable.
The heating medium flow rate has been varied in the range 20 to 60 lpm. It appears that, for
the range considered, changes in this flow have a limited impact in the COP values.
The impact of the heating medium temperature is shown in Figure 2.15. When the inlet
cooling water temperature is 30°C, the system starts showing reasonable performance only
above 100°C but the COP only approaches 0.60 for heating medium temperatures higher than
110°C.
0.700
0.600
charging concentration = 42.5%
0.500
COP [-]
0.400
0.300
0.200
0.100
charging concentration = 40.0%
0.000
80 90 100 110
Inlet hot water temperature [C]
Figure 2.15 COP of the system as function of the heating medium temperature.
0.700
0.500
0.400
COP [-] charging concentration = 40.0%
0.300
0.200
0.100
0.000
80 100 120 140 160
Weak solution mass flow rate [kg/h]
Figure 2.16 COP of the system as function of the weak solution mass flow rate.
Figure 2.16 shows that the charging concentration has a significant effect on the required
temperature of the heating medium to obtain acceptable performance values. Most probably a
further increase in concentration would have a positive effect.
The chilled water flow has been varied in the range 20 to 40 lpm. The impact on performance
was very small (2 to 3%). Also the chilled water temperature has been varied in the range 11
to 19°C showing an even smaller effect (less than 1%).
Finally the impact of the weak solution mass flow rate on the COP has also been investigated.
For the higher charging concentration, the COP is, in the tested range, independent of the
weak solution flow rate. The cooling capacity increases by 7% as the weak solution flow
increases from 80 to 160 kg/h.
Discussion
When the solar collector efficiency is combined with the experimental performance of the
NH3/H2O refrigeration system given in Figure 2.15 then the performance of the combined
system shown in Figure 2.17 results.
0.35
evacuated tube collector (Sydney type)
0.3
evacuated tube collector
0.25
COP_Solar [-]
flat-plate collector
0.2
stationary CPC collector
0.15
0.1
0.05
0
70 80 90 100 110 120
Average heating medium temperature [C]
Figure 2.17. Solar COP as a function of the average heating medium temperature
100
20
flat-plate collector
0
70 80 90 100 110 120
Average heating medium temperature [C]
The solar collector efficiencies apply for a radiation on the collector of 800 W/m2 and have
been taken from [Henning 2004]. The results shown in Figure 2.17 have been used to quantify
the required investment in solar collectors for a 10 kW refrigeration capacity. The specific
collector costs have been taken from [Henning 2004]. The flat-plate collector shows a
relatively low solar COP. Nevertheless it leads to the lowest solar collector investment. The
lowest investment level is obtained for average heating medium temperatures slightly above
100°C.
Although the flat-plate solar collector combines into the lowest investment level, it is still an
extra investment of about 1500 Euro / kW cooling capacity on top of the chiller investment.
Heat driven systems that allow lower driving temperatures would make solar sorption systems
more competitive.
2.2 Conclusions
Acceptable performance for the tested NH3/H2O single effect cycle was only obtained for
heating medium temperatures above 100°C.
For a solar radiation of 800 W/m2, the lowest solar collector investment level is obtained in
combination with a flat-plate collector operating with an average heating medium temperature
of 100°C.
Solar driven systems will become more competitive if systems requiring lower heating
medium temperature would be applied.
Half-effect H2O/LiBr chiller is better than single-effect NH3/H2O chiller for all cases. The
performance difference gets bigger with low-efficiency (i.e. low-cost) collectors.
originates from the low generator temperature and relatively high COP of HE H2O/LiBr
chiller.
Based on those results above, a half-effect H2O/LiBr cycle was thought the most adequate for
an air-cooled solar absorption cooling system.
Section 3.1 presents thermodynamic and transport properties of LiBr aqueous solutions in
wide working ranges.
Section 3.2 describes the thermodynamic design process and the results for the physical design
of a HE LiBr/H2O absorption chiller setup.
Section 3.3 includes physical details of the designs for all components and the overall system
configuration.
The last section explains the extra test setup designed for the investigation of heat and mass
transfer characteristics of falling film flows.
The contents in this section are quoted from authors’ theoretical study on the properties of
LiBr aqueous solutions. Only the result is presented here and more detailed information can be
found in Appendix A1.
The following equations have been derived for LiBr aqueous solutions in the concentration
range of 0 to 70 wt% and the temperature range of 0 to 210 °C:
where
T T
C p∞1
G1∞ = H1∞o − TS1∞o + ∫ C p∞1dT − T ∫ + V1∞ ( p − po* ) (3.2a)
To To
T
2 cj
C p∞1 = RT −2 ∑ (3.2b)
j =0 Tj
2 b0j
V1∞ = RT ∑ (3.2c)
j =0 Tj
T T
C lp 2
G2l = H 2l o − TS2l o + ∫ C lp 2 − T ∫ dT + V2l ( p − po* ) (3.3a)
To To
T
2
C l
p2 = R ∑ d jT j (3.3b)
j =0
2
V2l = R ∑ e jT j (3.3c)
j =0
G E = x1υ RT ⎡⎣ ln γ ± + (1 − φ ) ⎤⎦ (3.4a)
6
ibi
φ = 1 + ∑ (ai + p )mi / 2 (3.4b)
i =1 2υ
6
2 ib
ln γ ± = ∑ (1 + )( ai + i p )mi / 2 (3.4c)
i =1 i 2υ
⎣ {
p =exp ⎡α - ln (θ + θ 2 - 1) ⎤ /β
⎦ } (3.5a)
For maximum error of ±0.15 K in Tdew within –15~150°C from McNeely [1979]
1603.54 104095.5
log( p* ) = 7.05 − − (3.6a)
T T2
For maximum error of ±0.05 K in Tdew within –15 ~ 200°C from Schmidt [1979] and Perry et
al. [1999].
Details of the derivation and comparison with experimental results can be found in the
Appendix A1.
The calculated enthalpy is consistent with the experimental differential heat of dilution and
heat capacity data taken from the literature. This study agree relatively well with Feuerecker et
al. [1993] and Kaita [2001] but deviates substantially from those based on the equilibrium
vapor pressures of McNeely [1979] especially in the high concentration and temperature
regions.
Transport properties of working fluids were collected from various sources. To obtain fit
curves, in most cases, raw data or original correlations have been extended to fit the entire
concentration range.
The thermal conductivity of solution by Diguilio et al. [1990] has been extended down to pure
water as shown in the Figure 3.2 using Equation(3.7).
x10 -4
7.0
6.5
6.0
ksol (kW/mK)
0%
5.5
5.0
30%
4.5 40%
4.0
65%
3.5
250 300 350 400 450 500
Tsol (K)
Figure 3.2 Thermal conductivity of aqueous lithium bromide solution
3 4
ksol = ∑∑ a ij x j −1T i −1 (3.7)
i =1 j =1
where aij*1000=
i\j 1 2 3 4
1 -0.3863624126 -0.3122938151 17.75694663 -41.62113683
2 0.005245122201 -0.006413302194 -0.0800954908 0.2130478667
3 -6.398936707E-006 1.013622815E-005 0.00010029254 -0.000281450
The difference of Equation (3.7) from the original correlation of Diguilio et al. [1990] is
substantial for 30~40% solutions at high temperatures. The gradients of the original
correlation in the region were impossible to follow with a 3rd order polynomial function. It is
unclear if this trend is natural. Equation (3.7) is, however, safe to use for this project because
the working range of the system is far from the region of the large deviation.
The dynamic viscosity data from Diguilio et al. [1990] have also been used. The values from
the original correlation were correlated by Equation(3.8) between pure water and 65% as
shown in Figure 3.3.
4 4
ln μ sol = ∑∑ a ij x j −1T i −1 (3.8)
i =1 j =1
where aij=
i\j 1 2 3 4
1 15.4338601 -1.796143844 -453.964325 +1644.664107
2 -0.1496987184 +0.08581467986 +3.186981058 -11.18992719
3 0.0003210580467 -0.0004050019644 -0.006116119513 +0.02286554179
4 -2.397708795E-007 +6.025222928E-007 +2.699142889E-006 -1.3359444E-005
25
0%
10
1
250 300 350 400 450 500
Tsol (K)
Figure 3.3 Dynamic viscosity of aqueous lithium bromide solution
Surface tension is important for falling film flows. The falling film flows are more susceptible
to capillary forces when flow rates are small. Since the influence of surface tension on the
transfer characteristics of falling film flows were proven very strong, much interest has been
given to additives (surface active agents). But, no general theory is available for quantitative
description of additive’s influence and only incomplete experimental data are available.
100
60%
90
80 10%
0%
σ (mN/m)
70
60
50
40
30
250 300 350 400 450 500
Tsol (K)
Figure 3.4 Surface tension of aqueous lithium bromide solution
The surface tension data of pure solutions from Yao et al. [1991] have been curve-fitted by
Equation(3.9).
3 3
σ sol = ∑∑ a ij ( xw ×100) j −1T i −1 (3.9)
i =1 j =1
where aij=
i\j 1 2 3
1 100.0572325 -38.32451999 +72.41610217
2 - 0.02832353108 + 0.2991780643 - 0.3763508519
3 - 0.0002198127185 - 0.0004491657076 + 0.0009148617576
3
ln kvap = ∑ a iT i −1 (3.10)
i =1
x10 -6
40
35
kvap(kW/mK)
30
25
20
15
250 300 350 400 450 500
T(K)
Figure 3.5 Thermal conductivity of saturated steam
Viscosity data of saturated steam were also collected from Schmidt [1979].
x10 -6
16
14
μ vap(Pas)
12
10
8
250 300 350 400 450 500
T(K)
Figure 3.6 Dynamic viscosity of saturated steam
3
ln μ vap = ∑ a iT i −1 (3.11)
i =1
3.2.1 Introduction
Thermodynamic (cycle) design results for the proposed chiller are presented in this Section.
Section 3.2.2 describes the working principles of two different SPHE cycles.
Section 3.2.3 explains simplified thermodynamic models specially developed for the purpose
of cycle design.
SPHE cycles can be subdivided into two categories depending on the configuration of solution
flows. One is a series and the other is a parallel flow cycle. The working principles are
explained in the following.
Figure 3.7 shows a series cycle on a Dühring chart. The left-most line is the saturated steam
line where a condenser and two evaporators are located.
The cycle has three pressures, namely high-pressure Ph, mid-pressure Pm and low-pressure Pl.
A pressure can be calculated using a P-Tdew function, Psat(Tdew) from the given Tdew.
For example, in Figure3.7, CON (condenser, point 3) and GEN (generator, 16→24) pressure
can be calculated from T3, i.e. Ph=Psat(T3). In the same way, MPE (Mid-Pressure Evaporator,
point 2) and MPA (Mid-Pressure Absorber, 23→21) pressure is Pm=Psat(T2). The pressure of
LPE (Low-Pressure Evaporator, point 1) and LPA (Low-Pressure Absorber, 20→18) is
Pl=Psat(T1).
Process description is given below. Numerical indices for state points can be referred to either
Figure 3.7 or 3.8.
From the outlet of LPA (18), a solution pump lifts strong solution to generator (16→24)
through LT-Hex (Low-Temperature Heat EXchanger, 5→17) and HT-Hex (High-
Temperature Heat EXchanger, 17→16). In the generator, the strong solution is boiled by a
heating medium flow (35→37) fed from solar collectors and the steam is supplied to the
condenser (3).
The condensed refrigerant (3) is split in two and supplied to LPE (1) and MPE (2) and
eventually absorbed by LPA (20→18) and MPA (23→21) respectively. Among these two
split flows, only the refrigerant in LPE (1) brings a cooling effect. The other refrigerant in
MPE (2) cools down LPA (20→18), which enables the low-temperature generation (15→24)
of this cycle.
If the liquid refrigerant from the condenser can be cooled down somehow before being
supplied to the evaporators, it helps reducing the flash losses. It is possible to put such a heat
exchanger, namely Ref-Hex (Refrigerant Heat exchanger), between condenser and LPA so
that the refrigerant from the condenser is cooled down (3→4) while the cold strong solution
from LPA is warmed up (18→5).
The hot and concentrated weak solution at the outlet of generator (24) flows to MPA (23→21)
after giving up the heat (24→23) to the cold strong solution (17→16) in HT-Hex. The weak
solution is cooled by coolant flow (37→38) and absorbs the vapor from MPE (2) in MPA(
23→21).
The solution at the outlet of MPA (21), in turn, flows further into LPA after giving up the heat
(21→20) again to the cold strong solution on the other side in the LT-Hex (5→17). It absorbs
the vapor from LPE (1) while it is being cooled by MPE (2) and the solution finally returns to
the solution pump to make a complete cycle.
As already mentioned, the sole purpose of MPE (2) is to cool down LPA (20→18) at a below-
ambient temperature so that the generator temperature T24 can be significantly lowered.
Consequently, some refrigerant is sacrificed for the low generation temperature.
The other cycle, slightly different from the series cycle, is a parallel flow cycle. In a series
cycle, the weak solution from generator goes through two absorbers in series
(24→22→21→19→18). But in the parallel cycle, it is split after HT-Hex and the two split
flows go to MPA and LPA in parallel (24→22→21, 24→22→19→18). The two flows are
mixed again at the outlet of MPA.
This cycle is similar to the series cycle in performance but is more practical due to some
reasons to be explained later. Block and cycle diagrams are given in Figure 3.9 and 3.10.
3.2.3 Modeling
The simulation model introduced in this section was specially developed for the
thermodynamic design of SPHE cycles.
In the following sections, all numeric references of state points are based on the schematic
cycle diagrams shown in Figure 3.11.
3 15 24 3 15 24
16 16
33 34 36 35 33 34 36 35
4 4
14
2 23 2 23
17 17 21
21 22 22
37 38
37 38
1 5 20 1 5 20
18 19 18 19
32 31 32 31
(a) Series cycle (b) Parallel cycle
Figure 3.11 SPHE cycles
All transfer processes in a system were assumed heat transfer-dominated, i.e. mass transfer
resistance is neglected, and simplified so that they could be modeled by the ε-NTU method
[see e.g. Holman (1986)].
The biggest advantage of the ε-NTU model is that ε is linear with temperatures and thus all
transfer equations are easy to solve. A disadvantage is that it is difficult to take full account of
local variations of physical states within the components. Therefore, this approach is only
acceptable for the cases where variation of local states is not so important such as
thermodynamic design of a cycle.
Condenser
The vapor flow m& 03 from generator (16→24 in Figure 3.11) condenses in a condenser (3) of
size εCON at pressure p03=ps(T03). The coolant flow of m& 33 enters the condenser at T33 (oC).
Then the condenser can be represented by the following equations.
Mass balance:
where m& 01 and m& 02 are the refrigerant flow rates distributed to LPE (point 1) and MPE (point
2) respectively.
Heat balance:
where ΔTsup is the superheat of the vapor coming into the condenser, which is defined by
where Tv03 is the vapor temperature from generator and T03 is the dew temperature in the
condenser.
Heat transfer:
In order to save efforts to call a property subroutine whenever Equation (3.13) has to be
solved and also to avoid iterative solution, the latent heat of saturated steam, hfg has been
made a 2nd-order polynomial function of Tdew. For example, h03fg is given by
where a1=2495.62, a2=-2.0126 and a3=-0.0037682 can be used within maximum 0.5% error of
the steam data from Schmidt (1979) for 0 oC < Tdew < 212 oC.
Then Equation (3.13), (3.15) and (3.16) can be combined to give a solution for T03 as follows.
-b 2 + b 22 -4b1b3
T03 = (3.17)
2b1
where
b1= a3 m& 03
b2= (a2 + Cp03v) m& 03 + εCON m& 33C p 33
b3= -εCON m& 33C p 33 T33 - m& 03 (a1+ Cp03v Tv03)
Low-Pressure Evaporator
Heat balance:
∑ Q = m& 31C p 31 (T31 − T32 ) − m& 01 ⎡⎣ −C lp 01ΔT sub + h01fg + C lp 01ΔT sup ⎤⎦ = 0 (3.18)
where ΔTsup is the superheat of the vapor leaving the evaporator, which is defined by
and ΔTsub is the subcooling of the liquid entering the evaporator, which is defined by
Heat transfer:
Like the condenser, the dew temperature T01 could be solved as follows.
-b 2 + b 22 -4b1b3
T01 = (3.22)
2b1
where
b1= a3 m& 01
b2= (a2+ Cp01v-Cp01l) m& 01 + εLPE m& 31C p 31
b3= -εLPE m& 31C p 31 T31- m& 01 (a1+ Cp01v Tv01- Cp01l T04)
Generator
Solution flow with m& 15 at x15 and T16 is supplied to a generator of size εGEN at pressure
p03=ps(T03). Heating medium flow of m& 35 enters the generator at T35. Solution flow with m& 24 at
x24 and T24 leaves the generator. Superheated vapor flow of m& 03 leaves for condenser with
Tv03. Then the governing equations are as follows.
Heat balance:
∑ Q = m& 35 C p 35 (T36 − T35 ) − (m& 24 h24 + m& 03h03v − m& 15 h16 ) = 0 (3.25)
where hv03 is enthalpy of the superheated vapor that flows to the condenser and h16 is the
enthalpy of subcooled solution, which can be calculated by h16=h15-Cp15(T15-T16).
Heat transfer:
From Equation (3.23) and (3.24), m& 15 / m& 03 =x24/(x24-x15)≡α and m& 24 / m& 03 = α − 1 where α is the
ratio of (refrigerant-) rich solution flow rate per total refrigerant generated and is often called
‘circulation ratio’. Then, Equation (3.25) can be rewritten as follows:
∑ Q = m& 35 C p 35 (T36 − T35 ) − m& 03 ⎡⎣α ( h24 − h16 ) + h03v − h24 ⎤⎦ = 0 (3.28)
LPA/MPE unit
Refrigerant flow m& 02 from condenser evaporates at pressure p02=ps(T02) on the evaporation
side. Refrigerant-poor solution flow m& 19 enters absorption side (20→18 in Figure 3.11) at x19
and T20. The size of this component is represented by its effectiveness εLPA-MPE.
Heat balance:
where hv01 is enthalpy of the superheated vapor being absorbed and h20 is the enthalpy of
incoming solution, which can be calculated by h20=h19-Cp19(T19-T20).
Heat transfer:
From Equation (3.29) and (3.30), m& 18 = m& 01 x19/(x19-x18). And because m& 18 = m& 15 = m& 03α for
series cycle, Equation (3.31) can be rewritten as
⎡ m& ⎤
m& 03 ⎢α (h20 − h18 ) + 01 (h01v − h20 ) ⎥ − m& 02 (h02v − h04l ) = 0 (3.33)
⎣ m& 03 ⎦
⎡ ⎛x −x ⎞ ⎤
m& 01 ⎢α ⎜ 24 15 ⎟ (h20 − h18 ) + ( h01v − h20 ) ⎥ − m& 02 ( h02v − h02
l
)=0 (3.34)
x − x
⎣ ⎝ 24 18 ⎠ ⎦
Mid-Pressure Absorber
Poor solution flow of m& 22 at x22 and T23 is supplied to MPA (23→21 in Figure 3.11) with
effectiveness εMPA. Coolant flow of m& 37 enters the absorber at T37. Rich solution flow of m& 21
at x21 and T21 leaves the absorber. Superheated vapor flow of m& 02 at Tv02 from MPE is
absorbed.
Heat balance:
where hv02 is enthalpy of the superheated vapor from MPE and h23 is the enthalpy of the
incoming solution, which can be calculated by h23=h22-Cp22(T22-T23).
Heat transfer:
From Equation (3.35) and (3.36), m& 22 = m& 02 x21/(x22-x21). And because m& 22 = m& 24 = m& 03 (α − 1)
for series cycle from the mass balance equations of generator, Equation (3.37) can be rewritten
as
⎡ m& 02 v ⎤
∑ Q = m& 03 ⎢(α − 1)(h23 − h21 ) + & (h02 − h21 ) ⎥ − m& 37C p 37 (T38 − T37 ) = 0
m03
(3.40)
⎣ ⎦
⎡ ⎛ x24 − x15 ⎞ ⎤
∑ Q = m& 02 ⎢α ⎜ ⎟ (h23 − h21 ) + ( h02 − h23 ) ⎥ − m& 37C p 37 (T38 − T37 ) = 0
v
(3.41)
⎣ ⎝ x24 − x21 ⎠ ⎦
Condensed refrigerant flow m& 03 at T03 is cooled down to T04 on hot side of a heat exchanger
with effectiveness εRef-Hex. Cold rich solution flow of m& 18 at T18 enters the cold side and leaves
at T05.
⎛ x24 − x15 ⎞
∑ Q = m& 01α⎜ ⎟ C p18 (T05 − T18 ) − m& 03C p 03 (T03 − T04 ) = 0 (3.43)
⎝ x24 − x18 ⎠
Heat transfer:
In LT-Hex of size εLT-Hex, a refrigerant-rich solution flow m& 18 at T05 is heated to T17. A hot
refrigerant-poor solution flow m& 19 enters hot side at T21 and leaves at T20.
⎡ m& 01 ⎤
∑ Q = m& 03 ⎢α C p18 (T17 − T05 ) − (α − & )C p19 (T21 − T20 ) ⎥ = 0
m03
(3.45)
⎣ ⎦
Heat transfer:
In HT-Hex of effectiveness εHT-Hex, a refrigerant-rich solution flow m& 15 is heated to T16. A hot
refrigerant-poor solution flow m& 24 at T24 enters hot side and leaves at T23.
All simulations were focused on finding the best practical cycle for development of the
proposed chiller. First of all, all the relevant equations were normalized by m& 01. That is, in the
simulation program, all flow rates in the cycle are based on 1kg/s of the refrigerant used for
the effective cooling. Then a factor was multiplied to get results for design capacity.
0.40
8
9
0.38 11
8
α =13
COP
9
0.36
11
α =13
0.34
Series
Parallel
0.32
75 80 85 90 95 100
T24( C) o
Figure 3.12 Chiller COP vs. T24 for various circulation ratios
Figure 3.12 shows COP changes with the strong solution concentration, x18 for some
circulation ratios. The generator temperature T24 is, however, used on the x-axis instead of x18
to show what generator temperatures are achievable for each case. One can see that COP of a
chiller increases with decreasing generator temperature and with decreasing circulation ratio.
The reason is that the sensible heat loads in the generator and LPA both increase with x18 (T24)
and α.
32
28 α =13
Ahex(m2)
11
24 9
α =13
8
20 11
10kW cooling 9 8
Series
Parallel
16
75 80 85 90 95 100
T24( C) o
Figure 3.13 Total heat exchange area vs. T24 for various circulation ratios
Figure 3.13 shows the required heat exchange area for the same cases. Total heat exchange
area, Ahex is the sum of heat transfer areas of all components in a cycle.
Ahex increases exponentially approaching the minimum T24. It is the contribution of MPA
(A22_21). That is because, as T24 approaches the minimum, the logarithmic temperature
difference between the coolant (T37) and the weak solution (T21) in the MPA decreases
exponentially. It is clear that Ahex decreases with decreasing α. And less heat exchange area is
required for the parallel cycle.
The solar collector area was also calculated. The results are shown in Figure 3.14.
76
α =13
10kW cooling
Series
11
72 9
Parallel
8
Acol(m2)
68
64
60 α =13
11
9 8
56
75 80 85 90 95 100
T24( C) o
Figure 3.14 Solar collector (Collector I) area vs. T24 for various circulation ratios
Collector area in Figure 3.14 is based on a particular solar collector, Collector I and its
specifications are given in Table 3.2. The influence of T24 is clear in Figure 3.14. Total three
collectors, namely Collector I, II and III, were considered. The efficiency curves of those three
solar collectors are shown in Figure 3.15.
0.8
0.6
η
0.4
I
II
III
0.2
0
0 0.04 0.08 0.12 0.16 0.2
(Tavg-Tamb)/I
Figure 3.15 Efficiency curves of solar collectors
In order to compare different cycles in terms of total cost, the price per unit area for some
collectors and heat exchangers were assumed. The specifications of the collectors are given in
Table 3.2. The relative price was defined as the ratio of the collector price to the heat
exchanger price. For example, price of Collector III is assumed same as heat exchangers per
unit square meter in Table 3.2, which is assumed as 75 Euro/m2.
If we assume a solar collector operates at average temperature of 100°C under solar radiation
of 800 W/m2 and ambient temperature is 32°C, then the reduced temperature T* is 0.085. The
efficiencies of Collector I, II and III are 43%, 27% and 8% respectively.
Based on the assumed prices in Table 3.2, total costs (collector + heat exchanger) were
calculated and the results are given below. In Figure 3.16, it is shown that the minimum cost
achievable with parallel cycle would be similar to that of series. The influences of different
types of solar collectors on total cost are shown in Figure 3.17.
x75 x 75
190 170
10kWcooling
Series
α =13
180 Parallel 160
α =13
170 150
11 α =8
9
160 8 140
α =13
130 10kW cooling
150 Flat I
11
Flat II
9 8 Flat III
140 120
75 80 85 90 95 100 75 80 85 90 95 100
T24(oC) T24( C) o
Figure 3.16 Total cost for series and parallel Figure 3.17 Total cost for series cycle
cycle systems with Collector I. system with different collectors
The systems with parallel-cycle chillers are not shown but the trends are similar to those in
Figure 3.17. The systems with Collector III are cheaper in the low temperature range but the
required collector areas are the largest. The minimum required collector areas are roughly 55,
70 and 80 m2 for Collector I, II and III, respectively.
It is now understood how x18 and α change T24 and consequently the system cost for different
chillers. From the results above, it is clear that the concentration of strong solution x18 and the
circulation ratio α should be chosen to optimize T24 and thus the system cost. Minimum cost
exists roughly at α=9 and x18=50% for series cycle and α=14 and x18=47% for parallel cycle.
Therefore a system should be designed near these minimum-cost points.
The influences of the rest of parameters were also studied in a similar way. The influences of
effectiveness ε of the components on chiller’s COP can be seen in Figure 3.18.
0.42
LPA/MPE
Ref-Hex
CON
0.40
COP
HT-Hex
LT-Hex
0.38
0.36
0.3 0.4 0.5 0.6 0.7 0.8 0.9
ε
Figure 3.18 Chiller COP vs. ε of components (series, α=9, x18=49.5%)
The COP increases linearly with the ε of the two solution heat exchangers. The influences of
the other components are not so strong on COP.
68
HT-Hex
LT-Hex
64
Acol(m2) CON
60 Ref-Hex
LPA/MPE
56
0.4 0.5 0.6 0.7 0.8 0.9
ε
Figure 3.19 Collector area vs. ε of components (series, α=9, x18 =49.5%)
The required solar collector area also changes with the components’ ε. The influences are
clearly shown in Figure 3.19. The linearly decreasing curves of HT- and LT-Hex are the
results of the increasing COP curves in Figure 3.18. Ref-Hex and LPA/MPE unit also do not
change the collector area much. Condenser’s ε has, however, a strong influence on collector
area. This is because of its direct effect on the generator temperature. Small condenser
increases the generator pressure and thus the generator temperature. Consequently, the
average solar collector temperature increases and, as a result, the collector efficiency becomes
lower.
In the following, the sizes of all components required for the target cooling capacity are shown
in terms of heat transfer area. Heat transfer correlations in Appendix C were used for this
purpose.
The ε of condenser is set to 0.9 in Figure 3.20. The required area is about 2 m2 and is relatively
insensitive to T24 and circulation ratio. The surface area of a 30m-long Ф20 mm bare tube has
approximately 2 m2 of surface.
2.04 5.0
Series α =11
Parallel
α =13 9
2.02 4.0
11
8
A(m2)
A(m2)
9
2.00
3.0 α =13
8
11
1.98
α =13 2.0 9
8
Series
11 9 Parallel
1.96 8
1.0
75 80 85 90 95 100 75 80 85 90 95 100
T24( C) o T24( C) o
Figure 3.20 Condenser area vs. T24 Figure 3.21 Generator area vs. T24
The heat transfer area required for generator is given in Figure 3.21. ε is set to 0.7 for all
cases. Roughly a 3 m2 of surface may be used as a reference value for the design. For
comparison, the shell-and-tube heat exchanger of Figure2.10 in Ch.2 has 2.76 m2 of heat
transfer surface.
8 10
Series 8
Parallel 9
11
6
α =13 α =13
A(m2)
A(m2)
8
4 11 1
9 9
α =13
8
11
2 11 α =13
9
8
Series
Parallel
0 0.1
75 80 85 90 95 100 75 80 85 90 95 100
T24( C) o T24( C) o
Figure 3.22 LPA/MPE vs. T24 Figure 3.23 MPA vs. T24
Figure 3.22 shows the required area for LPA/MPE unit. ε is set to 0.5 for all cases. Roughly
5m2 is needed for series cycle while 2 m2 is acceptable for parallel cycle. Heat transfer area of
5 m2 is extremely large in comparison with the other components. A 1m-long, Ф350mm
cylinder has a surface of about 1.1m2. If the cylinder can have an extended surface ratio of 2,
heat exchange area of 2 m2 is feasible within the dimension. But since MPE has a very small
flow rate on refrigerant side, complete wetting of the heat transfer surface would be difficult.
Figure 3.23 shows the required area for MPA. Note that y-axis is in logarithmic scale. As T24
approaches its minimum value, required heat exchange area increases exponentially. It is
shown that parallel cycle requires more area than series cycle for the same T24. 2 m2 of heat
transfer surface may be also reasonable for design of this component, which makes MPA
similar to condenser and LPA/MPE unit in terms of the required heat transfer area.
Since the solution heat exchangers, HT-Hex and LT-Hex, have significant influences on COP
of a chiller as shown in Figure 3.18, it is very important to achieve the highest possible
effectiveness for those components. It may seem relatively simple to design these single-phase
heat exchangers but it is not. High velocity flows for turbulence heat transfer may not be
possible due to the excessive pressure drop. For this reason, velocity of 0.3 m/s was used
within all flow channels of the heat exchangers. Consequently, the overall heat transfer
coefficients were smaller than 500 W/m2K.
The required heat transfer areas for HT-Hex with ε=0.7 are shown in Figure 3.24. LT-Hex
with ε=0.8 is shown in Figure 3.25. 1 m2 of heat transfer area seems reasonable for both
components.
1.6 3
Series
Parallel
1.4 2.5
α =13
α =13
1.2 2
11
A(m2)
A(m2)
11
1 1.5 9
8
9
0.8 1
α =13
8 11
0.6 0.5 9
Series
8
Parallel
0.4 0
75 80 85 90 95 100 75 80 85 90 95 100
T24( C) o T24( C) o
Figure 3.24 HT-Hex vs. T24 Figure 3.25 LT-Hex vs. T24
Judging from all aspects including the required collector area, heat exchange area and possible
system configurations, a parallel SPHE cycle has been chosen for the chiller development.
This decision will be further justified in the following sections. Specifications of the chosen
cycle are summarized in the Table 3.3.
I II III Remarks
COP chiller 0.38 Collector efficiency, η (%) 53.8 41 32 1. I=800 W/m2
Collector area, Acol (m2) 61.8 81.2 104 2. Tamb=32°C
3. η is based on gross dimension.
Cooling eff. (COP× η,%) 20.4 15.6 12.2
Many different physical designs can be considered for realization of the proposed chiller. But
there are some common rules in the design of a H2O/LiBr chiller. Some of them are briefly
introduced below.
The chosen cycle has three different sub-atmospheric pressures. They are typically, 12 kPa
for generator and condenser, 2 kPa for mid-pressure absorber and evaporator, and 1 kPa for
low-pressure evaporator and absorber. To have a sense of how low these pressures are, one
can recall that 1 kPa is equivalent to 100 mm of H2O column. Under these low pressures,
saturation temperature can dramatically change due to a small pressure difference. One
example is given in Figure 3.26.
80
60 50%
Tsat(oC)
40
0%
20
0
0 2 4 6 8 10 12
P(kPa)
Figure 3.26 Saturation temperature vs. pressure
Since the system works in vacuum environment, the existence of inert gases inside the
system significantly deteriorates the system performance. Inert gas acts as a barrier for mass
transfer between transfer surfaces. The major source for such a gas is the corrosion of
construction materials. It is especially serious when there are mechanical or welded joints of
different metals immersed in high-temperature solutions. They become galvanic cells that
produce hydrogen. Although the maximum temperature of the proposed cycle is relatively
low, enough care should be given to selection of construction materials and manufacturing
methods. Copper and copper alloys were chosen for all transfer surfaces because of their low
corrosion rates in H2O/LiBr environment. A typical corrosion test result is shown in Figure
3.27.
Figure 3.27 Corrosion rates in a boiling aqueous 65% LiBr solution [Behrens, 1988]
The setup has to be as small and compact as possible while it is still easily accessible for
possible repairs and maintenances. It may sound contradictory to the previously mentioned
minimum pressure drop requirement. But, compactness does not necessarily mean large
pressure drops.
Components should be arranged in such a way as one can easily access any component
whenever it is necessary. Since all components should be developed simultaneously, it is
decided that all components should be easily separable so that they can be replaced or
modified with minimum time and effort.
Since tubing has to be avoided for vapor flows, the components having vapor flows between
them should be located in the same confined space. That means, generator shares a confined
space with condenser, evaporator with absorber. For example, a single-effect LiBr chiller
would have two confined spaces at different pressures connected by liquid tubing. It is
actually the way most of LiBr chillers are built.
For the proposed chiller, it would be economical and compact to provide those confined
spaces with vertical concentric cylinders with partition walls. Since it has three pressure
sections, three vertical concentric cylinders would provide three confined spaces, i.e. one
central space and two annuli.
To minimize heat loss from generator to ambient, it would be better to place a generator in
the center. And then the logical place for a condenser would be above the generator for small
diameter of the central cylinder. Following the magnitude of working pressures, the mid-
pressure components could be placed in the 1st annulus and the low-pressure components in
the 2nd annulus. The resulting 1st configuration of a series cycle machine is shown in Figure
3.28. A parallel cycle configuration is different from the series one only in solution tubing.
This is one of the early conceptual designs. As mentioned above, three cylinders form three
different pressure sections. The generator is a vertical falling film type heat exchanger and all
the other components are horizontal tubular type except the two solution-heat exchangers
(LT-Hex & HT-Hex). This early concept has a heat transfer loop with a circulation pump
between LPA and MPE. The disadvantage of the concept is that the secondary loop itself is a
big resistance in heat transfer causing large temperature difference between LPA and MPE.
The heat of absorption in LPA should be transferred to MPE in such a way that the
temperature difference between them is minimized so that the generator temperature is as
low as possible. The heat should be transferred directly through the wall of Cylinder 2. So,
Cylinder 2 should play the roles of partitioning and transferring heat at the same time. This
idea is shown in Figure 3.29.
Figure 3.29 Configuration II (Series cycle) Figure 3.30 Configuration III (Parallel cycle)
Now the LPA and MPE became falling film flows on each side of a vertical wall. This idea
ensures not only the minimum temperature difference but also provides a much simpler
construction.
The series cycle has, unfortunately, a serious problem with hydraulics of a solution flow in
this configuration. In the series cycle, the solution at the bottom of mid-pressure section
should be supplied to the top of the LPA. If we assume a difference of height ΔL from the
bottom of mid-pressure section and the top of LPA, the pressure difference between Pm and
Pl should be larger than ρsolgΔL, i.e. ΔL<(Pm-Pl)/ρsolg. But in all cases, it turned out that Pm
was not high enough and (Pm-Pl)/ρsolg was less than 100 mm. So, an extra pump would be
necessary to lift the solution to the top of LPA for the series cycle.
On the other hand, parallel cycle has no such problem. It is because the solution flow to LPA
is pushed by Ph not by Pm. This is one of the reasons why parallel cycle has been chosen for
development. The parallel configuration is shown in Figure 3.30.
3.3.3 Generator
In order to provide maximum heat exchange area and minimum vapor-side pressure drop
within the configuration given in Figure 3.30, original design was made in a shell-and-plate
type as shown in Figure 3.31a.
Φ220mm
Solution in
50mm
1000mm
44 tubes Heating
medium in
Solution out
The Shell-and-Plate generator in Figure 3.31a was designed to have 3.7m2 of heat transfer
surface and 2.6x10-2m2 of vapor flow area within a 1m-long Φ200mm shell. On the other
hand, the Shell-and-Tube generator in Figure3.31b (the same generator used for ammonia
absorption chiller in Figure 2.10) has 2.76m2 surface and 1.76x10-2m2 of vapor flow area in
the same shell. It is clear that a Shell-and-Plate type generator has many advantages over its
Shell-and-tube counterpart.
But the original design had problems in manufacturing. The Shell-and-Plate generator had 10
brazed plate heat exchangers with a solution distributor mounted to the top of each plate. The
most expensive part in the design was the brazing process, which takes more than 50% of the
estimated cost.
The brazing caused another problem. When a single generator plate was fabricated for trial, it
revealed unfavorable aspect of the design. A connector that was supposed to distribute the
heating medium to the flow channel between two outer plates has been completely blocked
by the soldering material (see Figure 3.32). Although such a design flaw could have been
corrected easily, the whole design concept for generator had to be rejected because of its high
cost.
Passage blocked by
soldering material Connectors
Center plate
Side plates
After considering several alternative generator designs, it was concluded that the project
could not be carried out within the budget if any generator were to be purchased. For this
reason, it was decided to reuse the Shell-and-Tube generator in Figure 3.31b. By this
decision, the heat transfer area is reduced to 74% of the original designs.
Figure 3.33 shows a schematic diagram of the final generator design. There are two
distribution devices for even distribution of the strong solution to the heat exchange tubes.
The first one is a circular tray with holes at the bottom. It holds the solution supplied from
above and distributes it around the outer perimeter of the tube bank. The purpose of this 1st
distributor is to minimize disturbances in the solution entering the tubes through the 2nd
distributor. The 2nd distributor is basically a piece of tube with three 2mm-wide 15mm-long
longitudinal slots inserted in each heat exchange tube. It was designed to introduce the
solution through the narrow slots into the tube. It also has two ring-shaped groves inside
intended to stretch the initial three streams of flow into a continuous film.
Φ220mm Solution
supply
Solution in
Holed tray
Slot
Distributor
50mm
1000mm
13mm
Baffle
Liq Vap
Tube
17mm
Solution out
Since solution and vapor flows in countercurrent in the generator, there is a risk of flooding.
The risk of flooding is highest at the narrowest section of the 2nd distributor, whose diameter
is 13mm. According to the original generator design, it has to generate 33 kg/hr of refrigerant
vapor from 395.3 kg/hr of 49.7% solution at 13 kPa. This means a vapor velocity of 19.5 m/s
in the 2nd distributor exceeding a flooding velocity of 14.7 m/s predicted by Diehl and
Koppany [1969] given by:
V f = F1 F2 (σ / ρ g )0.5 (3.52)
where F1=[di/(σ/80)]0.5 when di/(σ/80)<1 or F1=1.0 when di/(σ/80)≥1 and F2=( m& l / m& g )-0.25.
Once flooding takes place, it could be detected by oscillation of pressure. And the maximum
capacity of generator would be limited by the capacity at the moment. Because entrainment
of solution to condenser is also possible as it happens, the maximum capacity has to be
approached with a caution. If any amount of solution gets into the condenser, which may be
indicated by an increase in evaporator temperature, the entire refrigerant circuit should be
emptied and refilled with pure refrigerant.
Considering the predicted flooding velocity, a stable operation may be possible up to 70% of
the original design capacity.
3.3.4 Condenser
Because the original generator design had to be abandoned, that of condenser also had to be
changed. Figure 3.33 shows both the original and final condenser designs.
Φ400
Φ 500
Coolant out
Φ21, 24 turns
50
1050
in
25 Φ290
Φ220
Condenser had been redesigned in Shell-and-Coil type as shown in Figure 3.33b. The coil is
made from Φ21mm stainless steel tube. The coiling diameter is 290mm and the number of
turn is 24. Three vertical supporters secure the coil alignment. The condenser is mounted at
the top of the generator. Vapor condenses on the surface of the coil and flows down to the
receiver at the bottom.
A schematic diagram of the final design is given in Figure 3.34. Two distributors are
mounted at the top of the Φ400 copper tube on both sides. Liquids will be supplied through
the holes of the distributors at the top and collected though the tubes at the bottom.
Original design had a extended surface that would give 2.6 times more surface than the base
area. But this idea was also abandoned for financial reason. Final design has heat transfer
area of 0.94 m2, roughly 42% of the original design.
The solution distributor on LPA side is a circular box of a rectangular cross section with 62
evenly spaced Φ1.2mm holes at the inner wall. Since there is a small gap between the wall
and the LPA surface, which is intended to promote formation of a continuous film, the
solution is supposed to be stretched sideways first to form a film and then flows down over
the surface.
Refrigerant in
Solution in
Bare
10 Φ1 holes
MPE
62 Φ1.2 holes copper
surface
750
LPA
3
Solution 31 Φ1 holes
Refrigerant Φ400
distributor
distributor
Sol. out
Ref. out
Figure 3.34 Final LPA/MPE design
The refrigerant distributor on MPE side is same except that it has another distributor inside.
The distributor inside is a ring-shaped tube with evenly spaced holes at the top. The outside
distributor has 31 and the inside one has 10 Φ1mm holes. Such a double-distributor design
was necessary because the design flow rate is so small (17.8 kg/hr) that the flow distribution
would be seriously influenced by small disturbances. Also it was designed to accommodate
varying flow rates up to five times the design flow rate because complete wetting of the bare
heat transfer surface was thought difficult with the design flow rate and recirculation of
refrigerant is planned.
MPA is basically a coil made from bare stainless steel tubes with a distributor mounted
above. The construction is rather simple. The distributor is a single ring-shaped tube with 63
Φ1.2mm holes at the bottom. Several tube supporters secure the alignment of tube coil as
shown in Figure 3.35.
Φ300mm
Distributor
540mm-19 turns
63 Φ1.2 holes
m
1. 3m
Supporter Φ2
Solution receiver
Spot
out Solution out welding
Cooling wat. in Solution in
A receiver at the bottom will collect the strong solution from the last turn of the coil.
Originally extended surface was also planned for this component but it was abandoned for
financial reason. Total heat transfer area is estimated 1.2 m2, which is roughly 54% of the
original design.
The construction is basically same as MPA. But the distributor is double-distributor design. It
was designed so for the same reason as in the case of MPE. A difference in the distributor
design is that the outer distributor has a circular cross section and small metal caps are
mounted over the holes at the top to convert high-velocity upward jets into downward
streams when the flow rate is high in case of refrigerant recirculation.
Extended surface was abandoned for financial reason and the total surface area of the final
design is 2.25 m2, which is 48% of the original design.
44 Φ1 holes
Cap
10 Φ1 holes
Φ510mm
Ref. in
Chilled
water out
600mm-21 turns
mm
1.3
Chilled Φ2
water in
Ref. receiver
Refrigerant out
There are two heat exchangers for solution and one for refrigerant in the system. Three heat
exchangers have been purchased from the heat exchanger company Cetetherm, which closely
fit the original design conditions. The plate heat exchangers are selected as follows.
The relative positions and the flow configurations between components should be carefully
decided because several separate flows are exposed to different system pressures and they are
sensitive to pressure drops in the flow passages.
There are three major flows in the system, namely strong, weak and refrigerant flows.
The strong (refrigerant-rich) solution is a mixture of the two independent flows from the
bottom of LPA and MPA (point 18 and 21 in Figure 3.37) and should be pumped back to
generator by a solution circulation pump (G1). The solution is exposed to two different
pressures in the two absorbers. And the solution from LPA should go through two heat
exchangers, i.e. the Refrigerant and Low-Temperature Heat Exchanger (Ref-Hex and LT-
Hex), before being mixed in a solution tank (Sol-tank).
E1 E2 32
SG1 E1 E2
33
Q2 φ6
Shut-off Needle
valve valve
SG2
φ21
31
36 φ6
φ21
φ16 ~ 21 7
18
37 38
21 6
3 M2
φ6
Q1
5 SG3
35 17 LT-Hex Ref- 4
Hex G2
Q3 SG4 U-Tube Ref-tank
sv6 E3
24 23 20
SG5 M4
sv2 E6
16 25
M1 G1
HT-Hex sv5 Sol-tank
Considering the pressure drops in the heat exchangers and the two different pressures the
solution is exposed to, relative positions of the related components and the tubing were
configured as in Figure 3.38.
The U-shaped solution tank (Sol-tank) should be high enough to counterbalance the pressure
difference in the two absorbers and large enough to accommodate an amount of solution
needed to cope with varying working conditions.
PL PM
500
φ100 700
400
φ21
200 φ6 200
70
Service valve Sol-tank
Ground
Ref-HEX LT-HEX
Figure 3.38 Strong solution tubing
The refrigerant flows between condenser and two evaporators are exposed to three system
pressures. In order to separate the high system pressure in condenser, a U-trap will be used
after the condenser. The refrigerant flows from the condenser and the two evaporators are
mixed in a tank (Ref-tank in Figure 3.39) and pumped to the top of the evaporators.
The U-shaped refrigerant tank (Sol-tank) should also be high enough to counterbalance the
pressure difference in the two evaporators and large enough to accommodate an amount of
refrigerant needed to cope with varying working conditions
PH
From condenser
PL PM
1600
40
φ100
φ40
500
450
400
φ21
G2
Ref-tank 200
φ6 U-Tube Service valve
Ground
Ref-HEX
Weak solution from generator is split into two flows after being cooled down in the High-
Temperature Heat Exchanger (HT-Hex). One flow is directly supplied to MPA and the other
to LPA after being further cooled down in Low-Temperature Heat Exchanger (LT-Hex).
There is no reservoir or mixing tank for this solution. Properly selected flow control valves
will be installed in both of the lines as shown in Figure 3.37.
The system has been designed to operate in wide ranges of working conditions. A
temperature-controlled electric heater will provide heating medium temperature up to 120°C,
which is higher than the design temperature by 30 K.
A frequency-controlled gear pump will be used for circulation of solution, which is capable
of delivering more than three times the nominal design flow rate. Another gear pump will be
for refrigerant, which can provide a refrigerant recirculation ratio up to 6.
Temperatures of all liquid flows, three system pressures and three major flow rates with
densities will be measured.
In the experiments, every controllable parameter will be varied one at a time. The results
will be closely analyzed and used to optimize the system.
At the top of each component, a service valve is provided so that any inert gas can be
removed. The solution and refrigerant mixing tank in the system are connected to each other
with a shut-off valve in the middle. If refrigerant is contaminated by solution, the shut-off
valve will be opened to completely empty the refrigerant tank. After the tank is emptied, the
valve will be closed and the system will be run to refill the refrigerant tank.
Main purpose of this experiment is to provide experimental data for modeling falling film
flows on vertical plate and to evaluate performance of different heat transfer surfaces.
The initial distribution of liquid over transfer surface is important for a falling film flow with
a small flow rate. There is a characteristic flow rate, called ‘MWR (Minimum Wetting
Rate)’, below which complete wetting of surface is impossible. A liquid film breaks up when
the balance to maintain the film is broken and it turns into streaks of flows wetting only a
small portion of the surface. MWR is a function of several parameters including the
Reynolds number, surface tension, contact angle and so on.
Hartely and Murgatroyd [1964] derived MWR (Γmin) correlations for laminar film flows on
vertical walls based on the balance of forces. His correlation is given in Equation (4.1).
0.2
⎛μ⎞
[σ (1 − cosθ )]
0.6
Γ min = 1.69 ⎜ ⎟ (4.1)
⎝g⎠
where θ is a contact angle that may in turn be determined by Equation (4.2) from Carey
[1992].
⎡ ⎛σ ⎞
0.5
⎤
θ = cos ⎢ 2 ⎜ sg
−1
⎟⎟ − 1⎥ (4.2)
⎢ ⎜⎝ σ lg ⎠ ⎥
⎣ ⎦
where σsg and σlg are surface tensions on solid-gas and liquid-gas interfaces. According to
Hartely and Murgatroyd [1964], the contact angle of water at 30°C on 1-inch copper tube
wall is 20°. Carey [1992] gave also the same value for water on typical surfaces of heat
transfer equipment. Inserting σlg=71mN/m for water at 30°C and θ=20° in Equation (4.2)
yields σsg=67 mN/m.
From Equation (4.1), Γmin is 35 kg/mhr for an adiabatic isothermal water flow at 30°C on a
smooth copper surface. Assuming the same σsg=67 mN/m, Γmin were estimated using
Equation (4.1) for all components in the system.
Although the estimated MWR values are only approximations, one can see most of
components will have to operate not far from the minimum flow rates. Especially in MPE
and LPE, serious wetting problems are expected. A possible solution is to use a surface
and/or chemicals that would promote wetting. This will be investigated using the proposed
film flow test setup.
When flow rate is large, even if solution is distributed in such a way that a number of
individual streaks of flows are formed at the beginning, they tend to merge into one to form a
complete film. But this is not easily done if the initial separate flows are not closely located.
From this reason, a good distributor has as many distribution points (holes, tubes and etc.) as
possible. An extreme example of such a kind may be the ‘weir distributor’ that let the liquid
overflow the weir evenly along the entire edge to form a complete film from the beginning.
But the weir design is not applicable when flow rate is small because surface tension of
liquid tends to act against the ‘even overflow’.
Let alone the ‘complete wetting of surface’, the ‘even distribution’ will be important
especially in the design of MPE and LPE. A distributor model had been developed and all
distributors in the chiller setup were designed using this model. Details of the model are
given in Appendix B.
There are some commercially available heat transfer surfaces for high heat and mass transfer
performance. Some examples from Fujita [1993] are given in Figure 4.1.
As is shown in the figure, the heat and mass transfer coefficients of the extended surfaces are
more than twice as large as the area of bare tubes. The enhancement in small flow rate region
(below MWR) is mainly due to the increase of wet surface ratio. In the large flow rate region
(above MWR), the mixing effect contributes the enhancement. Roughly speaking, 100~300%
of enhancement seems feasible.
Recent Kim et al. [2002] confirmed the prospect. They studied the influence of surfaces and
a surfactant on absorber’s performance and presented some interesting results given in Figure
4.2.
Figure 4.2 Effects of surface geometry and surfactant on heat transfer [Kim et al. 2002]
Although the surfaces used in the study do not show as much enhancement as Fujita’s
[1993], it is very interesting that the hydrophilic tube shows the highest heat transfer
coefficient. If an extended surface can have such hydrophilic characteristics, the
enhancement would be much greater.
Separation HX : precooler
tank
out Cooling water out
T4
FM2 HX Cooling Solution in T7 Cooling
700
water in water in
T5 T6
P Vapor in FM3
Distributor
Glass tube
T1
FM1
The generator has four 1.5 kW electric heating elements and one of them is temperature-
controlled, giving a continuous heat input range from 0 to 6.0 kW.
The separator is an empty pressure vessel with three connections. When a two-phase flow
from the generator enters the separator, the liquid is separated from the vapor and flows
down while the vapor goes out through the tube at the top. The flow rate of the separated
liquid is measured before it enters the absorber below.
The precooler right above the absorber cools down the incoming solution when it is
necessary.
The absorber consists of a copper plate heat exchanger in a 900mm-long φ150 mm glass
tube. The heat exchanger is attached to a steel lid that seals the glass tube at the top. The heat
exchanger consists of flange-tightened two thin copper plates with a rubber gasket in
between, on one side of which the solution is designed to flow in a thin film. In a rectangular
channel formed by the rubber gasket between the two copper plates, cooling water flows to
cool down the thin film flow outside absorbing vapor from the separator. The detailed design
of the heat exchanger is given in Figure 4.4.
φ10
40 φ6 5 5
25 1
20
φ6
540 550 20
20
15 1
15 6
95
Unlike single-phase heat transfer, absorption process involves somewhat complex phase
change phenomena at the vapor-liquid interface. This phase change at the interface provides
complex boundary conditions to heat and mass transfer. Determination of the interface
condition is important for the design and analysis of an absorption system in this aspect.
Because there is no device to measure any quantities at the interface in the present
experimental setup, determination of interface conditions will be done by heat and mass
transfer models. The basic concepts of the models are introduced below.
Because the local states cannot be measured, the results will be processed to obtain average
heat and mass transfer coefficients. Then, the first step is to define the average driving force
in heat and mass transfer.
y
z
xb mv
xi hv
Ti
qw dA
Tb qi
Tw Twall
In the control volume in Figure 4.5, heat flux at the wall can be expressed as
⎛ ∂T ⎞
qw = k ⎜ ⎟ ≡ hT ,bulk (T − T )
b wall
(4.3)
⎝ ∂y ⎠ y =0
The total heat rejection to coolant can be expressed in terms of overall heat transfer
coefficient and logarithmic mean temperature difference as follows.
(Tinb − Tout
w
) − (Tout
b
− Tinw )
Q& = ∫ U (T b − T w )dA ≡ U (4.4)
ln ⎡⎣(Tinb − Tout
w b
) /(Tout − Tinw ) ⎤⎦
Since the necessary bulk temperatures are measured to determine U in Equation (4.4), the
average film-side heat transfer coefficient hT ,bulk can be calculated back.
dm& v ⎛ ∂x ⎞
= ρ D ⎜ ⎟ ≡ ρ K (xb − xi ) (4.5)
dA ⎝ ∂y ⎠ y =δ
where K is the mass transfer coefficient based on bulk and interface concentration.
The total mass absorbed into the film can be expressed in terms of average mass transfer
coefficient and logarithmic mean concentration difference as follows.
Inaccuracy in determination of the K mainly comes from the difficulties in measuring the
interface concentrations, i.e. xiin and xiout. Although there are some techniques available for
measuring the interface concentrations, because experimental determination of the interface
concentration is beyond the scope, the models of Nakoryakov and Grigoreva (1980) and
Yüksel and Schlünder (1987) summarized in Ch. 8 will be used to predict the concentrations.
The models predict an interface concentration given the bulk concentration and temperature
at the corresponding position. Since the in- and outlet bulk concentrations and temperatures
are available from measurements, K can be determined.
A 2mm-thick plain copper plate was tested firstly in condensation mode, i.e. with only pure
water charged in the system. When the system operates, the steam from the separator
condenses on the copper plate and water from the separation tank will bypass the test section
(see Figure 4.3). The data were used for determination of the heat transfer coefficients on the
coolant side.
In the tests, coolant flow rate was varied while a fixed heating power was maintained. The
coolant flow was varied within 1.5-7.4 lpm, which is equivalent to Reynolds number range of
400-2200. During the tests, the coolant (tap water) inlet temperature was varied between 9.8-
13.9 oC.
Three 1.5 kW (nominal) heating elements were used in the tests. In total, three sets of tests
(1.5, 3 and 4.5 kW nominal capacity) were carried out and 41 data points were obtained.
Pressures (dew temperatures) and heat transfer rates measured during the tests are shown in
Figure 4.6.
1.5kW 1.5kW
80 3.0kW 6 3.0kW
4.5kW 4.5kW
4.5
Qcon(kW)
60
Tdew(oC)
3
40
1.5
20
400 800 1200 1600 2000 2400 400 800 1200 1600 2000 2400
Recool Recool
(a) Pressure vs. coolant Reynolds no. (b) Condensation heat vs. coolant Re no.
Figure 4.6 Pressure and heat transfer rate from condensation data
The heat transfer rate Q& con in Figure 4.6b was calculated from coolant side. As is shown in
the figure, Q& slightly decreases with decreasing Reynolds number, i.e. smaller coolant flow
con
rate. Since heat input is constant for each set of tests, the smaller Q& con for lower Reynolds
number can be understood as due to larger heat loss to ambient, which can be expected from
the higher dew temperatures in lower Reynolds number range in Figure 4.6a.
Whole absorber assembly has been insulated with averagely three layers of a 0.3 mm-thick
paper masking tape except for the test surface. Foam insulators could not be used because of
outgassing problem in vacuum environment. Although paper is a good insulator
(kpaper=0.05W/mK), condensation took place on its surface. The paper surface has been
observed wet during all the condensation tests and this has been taken into account in the
analysis.
Taking into account of the condensation on the paper surface, Q& con can be written as
where Q& front denotes the heat transferred through the test surface and Q& back through the rest of
surface, which is mostly the backside of the absorber.
Since Q& front transfers through a 2mm-thick copper plate to coolant flow, it can be written as
w
Tout − Tinw
ΔTcon ≡ (4.9)
ln ⎡⎣ (T dew − Tinw ) /(T dew − Tout
w
) ⎤⎦
And since Q& back transfers through firstly, three layers of paper tape, then 6mm-thick steel
plate to coolant flow, it can be written as
In order to determine the coolant-side heat transfer coefficient hcool, one should solve
Equation (4.7), (4.8) and (4.10) simultaneously. What makes this process even more complex
was that the condensation on the test surface was a combination of drop- and film-wise
condensation, for which no reliable correlation exists. On the other hand, condensation
pattern on the paper surface could not be clearly observed because the surface was always
wet.
Since heat transfer coefficient of drop-wise condensation generally reports dozens of times
higher than that of film-wise condensation and the condensation observed on the test surface
seemed somewhere between the two condensation regimes, any estimation based on
empirical correlations was thought highly unreliable.
An alternative is the Wilson plot method (Wilson, 1915). The idea is that experimental data
points obtained with different coolant flow rates are on a linear line in a so-called Wilson plot
and condensation heat transfer coefficient can be determined at intercept of the linear line
with the y-axis, if the heat transfer coefficient on the condensation side is kept constant
during the experiments. This is explained briefly in the following.
1 ⎛ AΔTcon ⎞ 1 Δt 1
⎜≡ & ⎟= + wall + (4.11)
U ⎝ Qcon ⎠ hT ,con kwall hT ,cool
If the condensation heat transfer coefficient hcon is kept constant, with 1/ U determined from
measured heat and temperatures, Equation (4.11) is a function of the coolant-side heat
transfer coefficient hcool only. hcool is described from a general form of Nusselt number as
⎛ k ⎞
hcool = C Rem Pr n ⎜ ⎟ (4.12)
⎝ dh ⎠
Inserting Equation (4.12) into Equation (4.11) and arranging gives a linear equation
1 dh ⎛ 1 ⎞ ⎛ 1 Δt ⎞
= ⎜ m n ⎟ + ⎜⎜ + wall ⎟⎟ (4.13)
U Ckcool ⎝ Re Pr ⎠ ⎝ hT ,con k wall ⎠
If hcon is constant, Equation (4.13) is a linear equation in the 1/ U -1/(RemPrn) plane. This
linear line is called Wilson plot.
⎛1⎞ 1 Δt
⎜ ⎟ = + wall (4.14)
⎝ U ⎠ Re→∞ hT ,con kwall
and therefore hcon can be determined from intercept of the linear line Equation (4.13) at y-
axis.
The Wilson plot method is useful when heat transfer coefficient should be determined for
heat exchangers of complex geometries or for which there is no correlation available, which
is the present case.
Application of the Wilson plot method in the present case is a little complicated due to the
presence of unintended condensation on the insulated part of the absorber. Nevertheless, the
overall heat transfer coefficient in the test section, U front can be written, analogously to
Equation (4.13), as
1 dh ⎛ 1 ⎞ ⎛ 1 Δtcopper ⎞
= ⎜ m n ⎟ + ⎜⎜ + ⎟⎟ (4.15)
U front Ckcool ⎝ Re Pr ⎠ ⎝ hT ,con , front kcopper ⎠
In the simultaneous solution of Equation (4.7), (4.8) and (4.10), Nusselt’s solution for film-
wise condensation has been used for hcon,back. This choice seemed natural considering that the
insulated surface was completely wet during the whole tests. But hcon,front has been
determined by iteration. Starting from an assumption, hcon,front was updated graphically from
Wilson plots using Equation (4.15) until it converged.
Another difficulty in the present application of Wilson method is that the condensation heat
transfer coefficients could not be kept constant during the tests, which can be expected from
Figure 4.6a. Because dew temperature strongly influences condensation heat transfer, it can
be expected that condensation heat transfer coefficient must have changed during each set of
the tests. Constant pressure during the tests could have been achieved if coolant inlet
temperature had been controlled. In order to minimize errors due to the deviation in dew
temperatures, only those data within Re=1200 to 2000 have been used. Data above Re=2000
have been rejected because the flows may be non-laminar.
Figure 4.7 shows Wilson plots with different exponents for each set of the tests.
0.5 0.5 0.5
0.6 0.6 0.6
0.5 0.5
0.4 0.5 0.4 0.4
0.4 0.4
0.4
0.3 0.3 0.3
1/Ufront
1/Ufront
1/Ufront
m=1/3 m=1/3
m=1/3
The correct exponent of Reynolds number would give best agreement between the coolant-
side heat transfer coefficients from those different sets of test results in Figure 4.7.
Firstly, note that intercept value increases from Figure 4.7a to Figure 4.7c for a constant
exponent. Since condensation heat transfer coefficient is inversely proportional to the
intercept, this means that condensation heat transfer coefficient decreases with increasing
capacity.
And secondly, in Figure 4.7a, the lines for exponents 0.4 and 1/3 have negative intercepts on
y-axis. This means that the exponent should be bigger than 0.4 so that condensation heat
transfer coefficient can be positive in all cases.
It turned out that smaller exponent gives better agreement between the coolant-side heat
transfer coefficients from the three sets of test results, which narrows it to be between 0.4 and
0.5. Considering the changing pressures during the tests and the consequent errors, the
exponent m was set to 0.45.
Figure 4.8 shows coolant-side heat transfer coefficients determined with m=0.45. In order to
take into account the property variation due to bulk and wall temperatures, the viscosity ratio
(μ/μw)0.14 has been introduced into the correlation.
20
16
12
Nu
1.5kW
4
3kW
4.5kW
0
0 20 40 60 80
Re0.45Pr1/3(μ /μ w)0.14
Figure 4.8 Coolant-side heat transfer coefficients (m=0.45)
0.14
⎛ μ ⎞
Nu = 0.241Re 0.45
Pr 1/ 3
⎜ ⎟ (4.16)
⎝ μw ⎠
where all properties are based on average bulk temperature except for μw that is based on
wall temperature. The results are compared with some empirical correlations in Figure 4.9.
20
Nu
10
0
0 1000 2000 3000
Recool
Figure 4.9 Test results and some empirical correlations
It turned out that the coolant-side heat transfer coefficient is much larger than predicted by
the laminar correlation of Sieder and Tate (1936) using hydraulic-diameter concept.
Equation (4.16) predicts heat transfer coefficients within ±9.5% of test results as illustrated in
Figure 4.10.
3
+9.5%
hcal(kW/m2K)
-9.5%
2
1 1.5kW
3kW
4.5kW
0
0 1 2 3 4
hexp(kW/m K) 2
After condensation tests have been finished, the system was drained and charged with
10.26kg of 50 wt% LiBr-water solution. Due to incomplete drainage, the concentration of the
solution inside the system was found slightly lower than the original charging concentration.
During the preliminary tests, the amount of solution charged was found too much because the
solution overflowed into the absorber through the vapor line at the top of the separation tank.
For this reason, 1.3 kg of solution has been extracted leaving 8.96 kg of the solution in the
system. It is also observed that wet area on the test surface changes with time and history of
the flow variation, i.e. the wet areas were different for the same flow rate when the flow rate
was reached from a larger flow rate or when it was reached from a smaller flow rate. For this
reason, from the main results, only the data were used that have been recorded during stable
operation preceded by a warming-up period during which the solution is circulated at a flow
rate larger than the desired value.
Tests were carried out changing the test surface and working fluid for different solution flow
rates. For all cases, the coolant flow rate and the outlet temperature of the generator were
kept constant. The test cases are summarized in Table 4.2.
The test cases from 1 to 8 in Table 4.2 were intended to determine the influence of different
surface geometries and the presence of octanol in the solution on the heat and mass transfer
during the absorption process.
Two different surfaces have been tested. One is the bare surface of copper plate and the other
is the same copper plate covered with a copper wire screen. The screen is a standard product
woven with 0.38mm copper wires having 22 meshes per inch. It was not mechanically
bonded to the surface. It was inserted between the steel flange and the copper plate in Figure
4.4 and tightly bolted while it was stretched sideways. But complete contact between the
screen and the plate was impossible. Some portion of the screened test section had void space
between the screen and the plate surface.
Case 1, 2, 5 and 6 were intended mainly for determination of heat transfer coefficient
between bulk solution and test surface of the absorber. The condition of the solution at the
inlet of absorber was not controlled, i.e. the precooler in Figure 4.3 was not in operation. For
this reason, condition of the solution was close to the saturated state.
On the other hand, cases 3, 4, 7 and 8 were done without running coolant to simulate an
adiabatic condition at the test surface. During these tests, the coolant channel was kept dry
and the ends of coolant tubes were taped to stop any possible ventilation of air. These tests
were intended especially for determination of solution-side mass transfer coefficient at the
solution-vapor interface of falling film flow. The precooler in Figure 4.3 was operated at its
maximum capacity to cool down the solution at the inlet of the absorber.
Finally, for Case 9 and 10, the system was open to atmosphere to prevent steam generation so
that only the sensible heat transfer in falling film flows of pure water could be investigated
with or without the presence of octanol.
For the cases where the absorber was cooled down by coolant flow, heat transfer rate was
calculated by
b
(Tout − Tinw ) − (Tinb − Tout
w
)
ΔT film ≡ (4.19)
ln ⎡⎣(Tout − Tin ) /(Tin − Tout ) ⎤⎦
b w b w
From Equation (4.17) and (4.18), the bulk heat transfer coefficient hT,bulk is given by
−1
⎛ A front ΔT film Δtcopper 1 ⎞
hT ,bulk =⎜ − − ⎟⎟ (4.20)
⎜ Q& film kcopper hT ,cool
⎝ ⎠
where m& sol ,out is measured by the mass flow meter, FM1 in Figure 4.3, after the absorber and
the inlet mass flow rate can be calculated by
where the volumetric flow rate V&sol ,in is measured by FM2 at the absorber inlet in Figure 4.3.
Note that the density of solution ρ is denoted as a function of temperature and concentration.
Because the density of the solution at the absorber inlet was not measured, ρ in Equation
(4.22) had to be determined indirectly.
Since the mass of LiBr is constant in the whole process, a mass conservation equation is
written as
b
m& sol ,out xout − m& sol ,in xinb = m& sol ,out xout
b
− ⎡⎣ ρ (Tinb , xinb ) V&sol ,in ⎤⎦ xinb = 0 (4.23)
where the solution concentration at the absorber outlet xbout could be determined from density
measurement at that point. Since the temperature of the solution at the inlet was measured
right before the volumetric flow meter (T4 in Figure 4.3), Equation (4.23) is a function of xbin
only. By solving Equation (4.23) for xbin using a proper density correlation, the density and
the mass flow rate at the absorber inlet were determined and thus the absorption rate could
also be calculated from Equation (4.21).
Alternative way to determine mass flow rate at the inlet is to solve a heat balance equation
over the absorber, which is given by
m& sol ,in hinl (Tinb , xinb ) + m& v h v (Tinv , p ) − m& sol ,out hout
l
(Toutb , xoutb ) − Q& film = 0 (4.24)
⎡ x b l b b ⎛ xout b
⎞ v ⎤ &
m& sol ,out ⎢ outb
hin ( Tin , xin ) + ⎜ 1 −
xinb
⎟ h − hout ⎥ − Q film = 0
l
(4.25)
⎣ xin ⎝ ⎠ ⎦
where the measured volumetric flow rate V&sol ,in and the density ρ (Tinb , xinb ) at the absorber
inlet do not appear. Equation (4.25) is also a function of xbin only because all the other terms
are either directly measured or can be calculated from the measured data. Using a proper
enthalpy correlation, Equation (4.25) can be easily solved.
Solutions of Equation (4.23) and (4.25) were used to crosscheck the solution flow rates
determined in these two different ways. For all cases, the thermodynamic correlations
presented in Ch.3 have been used.
The difference in the flow rates determined by the two different equations, Equation (4.23)
and (4.25), was found less than 6 % for the cases where the absorber was cooled down and
less than 2 % for the adiabatic cases. It was found that Equation (4.23) tends to give a smaller
m& sol ,in , i.e. larger m& v , than Equation (4.25).
Figure 4.11 shows the test results in terms of bulk film heat transfer coefficient against the
mass flow rates at the absorber inlet. The results for the 6 cases of water-cooled absorption
test are shown except for the 2 cases of adiabatic absorption test in Table 4.2.
hfilm(kW/m2K)
2
0
0 10 20 30
minlet(kg/hr)
Figure 4.11 Bulk film heat transfer coefficient
An unexpected result is found between single-phase heat transfer tests for water. First of all,
it can be seen that the heat transfer coefficient of pure water is several times greater that that
of water-octanol mixture. Secondly, while the heat transfer coefficient of pure water rapidly
increases linearly with the flow rate up to 10 kg/hr (Γ≈100kg/m hr), that of water-octanol
mixture is constant. For both cases, it was observed during the test that the heat transfer
surface was not completely wet below the flow rate.
0.12
Screen: LiBr solution+Octanol 100ppm
Bare : LiBr solution+Octanol 100ppm
Bare : water +Octanol 100 ppm
Screen: LiBr solution+no additive
Bare : LiBr solution+no additive
Bare : water +no Octanol
0.08
Nufilm/Pr1/2
0.04
20 40 60 80 100 120
Refilm
Figure 4.12 Film Nusselt number
In Figure 4.12, unlike the other fluids, Nusselt number of pure water increases linearly with
Reynolds number. The other fluids show more or less constant Nusselt numbers within the
range of Reynolds number.
The influence of octanol on the heat transfer of water is surprising. While the Nusselt number
increases with Reynolds number for pure water, it is constant for water-octanol mixture. It is
interesting that octanol actually suppresses heat transfer performance when mixed with
water. But it should be noted that the water-octanol mixture may behave differently on
different surfaces because surface tension is sensitive to surface condition and construction
material.
In the absorption tests with pure LiBr solution, it was found that the screened surface only
slightly increase the heat transfer (see □ and ∆ in Figure 4.12). It is thought that the screen
helps wetting the heat transfer surface to some degree.
On the other hand, the screen seems to have a negative influence on the absorption heat
transfer of LiBr solution when octanol is mixed (see ■ and ▲ in Figure 4.12). During the
absorption tests of LiBr solution with octanol, highly intensive activities were observed on
the surface of falling film, which must have enhanced the transfer process. The screen is
thought to have suppressed these activities.
Similar trends can be also seen in mass transfer. Figure 4.13 shows the adiabatic absorption
test results in terms of non-dimensional numbers.
5
Adiabatic test
Screen+additive(100ppm)
exp=0.82
Bare+additive(100ppm)
4
Screen exp=0.84
Bare
Sh/Sc1/2
3 exp=0.77
2 exp=0.81
0
0 20 40 60 80 100
Remean
Figure 4.13 Film Sherwood number
Similar to heat transfer case, the screen helped the mass transfer when octanol is not present.
Its influence is, however, remarkably strong. With the screen, Sherwood number was
doubled (□ and ∆ in Figure 4.13). But when octanol is mixed, Sherwood number is larger
with bare surface (■ and ▲ in Figure 4.13).
The test results can be represented by the following correlations for 40 < Ref < 110.
Equation (4.26)~(4.31) are the fitting curves of the bulk film Nusselt numbers in Figure 4.12.
Following equations are the fitting curves of the Sherwood numbers in Figure 4.13.
All the fluid properties were calculated with average of the inlet and outlet bulk conditions.
4.6 Conclusions
Heat and mass transfer in falling film flows have been investigated within the range of
Reynolds number from 40 to 110, where most of components in the proposed chiller were
designed to operate. All test results have been correlated with non-dimensional numbers.
Regardless of surface- and working fluid type, test surface was not completely wet within the
Reynolds number range.
Octanol enhances heat- and mass transfer by two and three times respectively.
The copper screen suppresses activities on film surface induced by octanol and thus should
not be used together.
Octanol should not be used for evaporator because it significantly lowers the heat transfer
coefficient when mixed with water.
sv1
P1
34 E5
22 E4
19
P2 2 P3 sv 1
4
φ21 sv3
SG1 E1 32 E
2
33
Q2 φ6
SG2
φ21 31
36
φ6
φ21
φ16 ~ 21 18 6 7
φ6 3 21 37 38 M2
Q1
35 SG3
17 HX2 5 HX3 4
G2
Q3 U-Tube3
SG4 U-Tube2
23 20 sv6 E3
SG5 M4
sv2 24 25 U-Tube1
M1 G1
16 HX1 sv5
Unlike the diagram, the condenser and generator column on the left of the diagram is actually
located at a much higher position than the absorber-evaporator shell on the right in order to
secure a hydraulic head large enough for the solution from the generator to flow through the
many heat exchangers and measuring equipments on the way to absorbers. Besides there are
several U-traps and reservoirs (tanks) in solution and refrigerant lines, which are not shown
in the diagram. In order to help understand the real configuration, several photos of the setup
are presented in Figure 5.2.
After the setup was assembled, small and big leaks have been detected and sealed during the
extended period of preparation. Unfortunately, however, one single leak that could not be
sealed properly was in the flange that connects the condenser and the generator. After several
attempts, it was concluded that the damaged surface of the flange would never be able to
provide a satisfactory level of vacuum tightness unless it was reprocessed or replaced.
Considering the schedule and budget to meet, the idea of sealing the leakage was abandoned.
Instead, it was agreed to evacuate the system more frequently during the tests.
2.4
p(kPa)
2
1.8
0 1 2 3 4
time(hr)
Figure 5.3 Leak test result (whole system)
Figure 5.3 shows one of the final leak test result. The pressure was measured for the system
left at room temperature after the system has been evacuated. The system pressure increased
at a rate of about 0.115 kPa/hr. When the system is evacuated once a day, early in the
morning for example, the system high pressure could be roughly 1 kPa higher than the right
value in the late afternoon, i.e. after 10 hours. But this leak would not directly influence mid-
and low-pressure parts of the system because the leaking part, i.e. the condenser and the
generator, would have been separated in pressure-wise, once working fluids occupied their
flow paths.
Since the system consists of many components working under different pressures, design and
arrangement of the components were given much care from the early phase of the
development. The following sections describe the problems encountered during preliminary
tests and the measures taken to solve them before the start of main tests.
In order to use minimal number of moving parts, no pumps were considered for the working
fluid flows from the high-pressure components, i.e. the condenser and the generator, leaving
hydraulic and system static pressure differentials as driving force for those flows. Since the
system pressures are equal at the moment of start-up, the hydraulic head between the
generator and the absorbers should be secured to initiate operation. In the real setup in Figure
5.2, although the top of condenser is about 4.5m high above ground, because the condenser
and the generator are connected vertically, the bottom of the generator and the top of
absorber is vertically only 500 mm apart. This means that, before the high system pressure
forms in the condenser and the generator, the flow resistance between the generator and the
absorbers must be small enough for the 500 mm of solution head to deliver the required
solution flow rates. For this reason, the components between the generator and the absorbers,
i.e. heat exchangers, flow meter, distributors and etc., were designed and selected for
minimal pressure drops. During the preliminary tests, it was confirmed that the maximum
solution and refrigerant flow rates between the generator and the two absorbers exceeded the
original design flow rates.
To help better wet the evaporator surfaces, the chiller setup has a gear pump to recirculate
refrigerant. Figure 5.4 shows the schematic diagram of the assembly.
LPE
MPE
G2
Ref. tank Recirculation pump
The recirculation pump in Figure 5.4 was confirmed to cavitate under all the working
conditions tested. For this reason, the input frequency could not be increased above 25 Hz
yielding only 60 to 70 lit/hr of refrigerant flow rate compared to the desired 200~300lit/hr.
Even then, considering that the gear pump was not selected and purchased for this specific
application, it would not have been a big problem to replace it, if there had been any
replacement. Given the schedule and the budget, ordering a replacement was not an option
and it was decided to carry on with it. But, not to mention the performance of LPE,
considering that MPE has such a wide perimeter of flat bare surface to wet, this lack of
refrigerant flow would substantially compromise the overall system performance.
To make matters worse, it was found that, at least, some of the refrigerant in LPE escaped
from its flow path to LPA side. That is, the refrigerant was wasted not contributing to the
cooling and on the other hand, the solution on LPA side became constantly diluted. After
some time, it was concluded that the poor design of LPE was the reason, which is illustrated
in Figure 5.5.
Cap
Support
welded to wall
LPE 1.3
m
Φ2
m
LPA
Main flow 0~6 mm
Tray
Escaping flow
Figure 5.5 Escaping refrigerant in LPE
Alignment of the LPE coil in vertical plane was intended by attaching it to three vertical
supports, which is in turn attached to the inside wall of the shell, as shown in Figure 5.5.
Since LPE is spirally coiled, some of the refrigerant flows along the tube and must meet the
support. When its flow path is blocked by the support, a portion of the flow would change its
direction towards the bracket welded to the wall and would further flow downwards along
the wall. Since the refrigerant tray down at the bottom does not completely contact inner wall
of the shell, there is a gap between the tray and the wall. According to the drawing used to
fabricate the tray, the gap ranges from 0 to 6 mm. Therefore it is likely that some of the flow
flowing along the wall can easily escape the tray into LPA side. It would have been possible
to prevent this from happening, if either the tray had been designed to firmly fit the inner
diameter of the shell or properly designed supports had been used for the alignment of the
coil. Since fixing this problem was not less consuming than making a new shell, no attempt
was made to stop the refrigerant from escaping. Instead, it was agreed to collect more data to
reduce uncertainties in data analysis by modifying the refrigerant tubing and adding some
extra equipments as shown in Figure 5.6.
Flow
meter 2
2 valves
+2 F.M.
Flow
meter 1
G2
The original assembly shown in Figure 5.4 had two identical capillary tubes to evenly
distribute the refrigerant to the two evaporators. In Figure 5.6, however, two control valves
and one more flow meter were included to control and measure the exact amount of flow
rates to both evaporators.
A problem found in the MPA could be solved relatively easily. It was found that the flow
resistance in the suction line of MPA was too large for the desired flow rate. For this reason,
an overflow took place from the MPA’s solution tray. From the MPA to the solution tank, the
solution was supposed to flow by the combination of pressure difference and the solution’s
hydraulic head as shown in Figure 5.7.
LPA MPA
m1 Overflow
m2
Resistance
P1
P2
G1
Solution tank
P1<<P2
Ref-Hex LT-Hex m1>>m2
Figure 5.7 Overflow from MPA
But it was found that the flow resistance in the solution line between MPA and the solution
tank was too large to accommodate the original design flow rate. In order to increase the
flow, instead of using the double-tank configuration as in Figure 5.7, it was decided to use
only one solution tank for both solution flows from the MPA and LPA as in Figure 5.8.
P1 P1
G1 G1
By modifying the tubing as in Figure 5.8, it was hoped that the strong suction pressure
generated by the decreasing level of solution in the single solution tank can effectively draw
the solution flows from both absorbers at the same time. The attempt was successful and no
overflow was observed under all working conditions.
Sign of the impurity of refrigerant did not disappear even after the overflowing problem of
MPA was gone. The density measurement from one of the flow meters in Figure 5.6
suggested the existence of heavy substances in the refrigerant after repeated blow-down
(drain and refill of the refrigerant) procedures. After some time, it was found that there was a
hot narrow vertical patch on the surface of the shell surrounding LPE, which is supposed to
be coldest in the whole setup. And by no accident, the patch was located right below the
solution supply tube to LPA as shown on the right in Figure 5.9.
Solution
supply tube 1 2
Solution
supply tube
LPE
LPA Hot
patch
Unlike the original design in Figure 5.9, the solution supply tube pass through the outer shell
to reach the LPA’s distributor, which was improvised at the time of assembly. It was
suspected that either the connection where the tube is welded to the outer shell (marked with
the circle #1 in Figure 5.9) or the hole of the tray that tube finally fits (marked with the circle
#2 in Figure 5.9) was not completely sealed. Anyway, solution must have escaped from
somewhere between point 1 and 2 in Figure 5.9 and have flowed down along the inner wall
of the shell forming the hot patch absorbing some of the vapor from LPE. And some of this
escaping solution must have fallen into the LPE’s tray in Figure 5.5 to contaminate the
refrigerant.
Locating the exact leaking point and sealing it was not an option at the moment because the
test plan was already delayed far behind the schedule. As a compromise to carrying out with
the impure refrigerant, in order to maintain the impurity of refrigerant below an acceptable
level, a ‘blow down valve’ was installed to the setup as in Figure 5.10.
G2 G1
By empting the refrigerant tank before operating the setup, the refrigerant contaminated
during the previous operation could be removed so that LiBr should not accumulate in the
refrigerant tank. The blow down process has been carried out whenever it was thought
necessary even while the setup was in operation.
After all necessary measures were taken to cope with the problems found in the preliminary
tests but before the originally planned tests were carried out in full scale, several attempts
were made to charge the setup with LiBr solution. The volume and concentration of charging
solution was changed several time until the setup ran properly, though not optimally, by trial
and error. The criteria for the ‘proper charging’ were firstly, the solution flows steadily at the
design flow rates and secondly, the concentrations should be reasonably close to the design
conditions. After several trials, the setup was ready with 28 kg of 50 LiBr %. It should be
mentioned that this is a rough estimation because the solution and refrigerant charged for the
preliminary tests could not be drained completely from the setup. But since the goal is to test
the performance of laboratory system, not to determine the optimum charging condition for a
product, it is not so important in the present phase of the project.
Octanol had been already added to the charged solution at 100 ppm of mass concentration
(3.35 ml in 28 kg). Although this first charging condition was by no means optimum, several
sets of tests were carried out as summarized in Table 5.1.
Table 5.1 Parameters and the test values (28 kg of 50% solution + 100ppm octanol)
Parameters Step1 Step2 Step3 Step4 Step5 Step6 Remark
T31 (oC) 13 15 20 24 chilled water temp.
T35 (oC) 65 70 75 79 84 89 Heating medium temp.
T33 (oC) 33 37 40 coolant temp.
m35 (lpm) 20 25 30 35 40 heating medium flow rate
m33 (lpm) 20 25 30 35 40 chilled water flow rate
m31 (lpm) 15 17.5 20 22.5 25 coolant flow rate
m01 (kg/hr) 12 17 22 27 32 37 ref. flow rate to LPE
m19 (kg/hr) 130 140 150 160 170 180 sol. flow rate to LPA
1. Standard values are in the gray cells, to which the other variables are set while a variable is changed.
2. Numerical indices should be referred to Figure 5.1
3. Step values other than standard ones are nominal. The real measured value can be different
4. No standard value was set for m19. Instead, the solution from the generator is distributed to two absorbers as
evenly as possible.
Tests were carried out to determine the influences of chosen parameters on the system
performance. Standard values (in gray cells in Table 5.1) were chosen for the parameters as
close to the corresponding design conditions as possible. For each series of the tests, only the
single parameter of interest was changed according to the steps in Table 5.1 while the other
parameters are set to the corresponding standard values. Among the standard values for the
parameters, T31, i.e. the chilled water temperature at the evaporator (LPE)’s inlet, is set to
24°C, exceptionally far from the original design value of 12.5°C. This is to compensate the
deteriorated system performance due to the impure refrigerant in evaporators, which will be
explained later in details.
In the following sections, test results are presented for each parameter.
In order to explain why the standard value for T31 in Table 5.1 was set to 24°C, first of all,
the influence of the chilled water temperature on the cooling capacity is shown below.
4
3
QLPE(kW)
0
12 16 20 24 28
T31( C)
o
Figure 5.11 shows the change of cooling capacity against chilled water temperature. The
cooling capacity, QLPE sharply decreases from 3.3 to 0.7 kW while the chilled water
temperature, T31 decreases from 24.5 to 13°C. Such a small cooling capacity around the
original design temperature (12.5°C) forced the use of a higher chilled water temperature
because otherwise, errors of measurement devices would seriously harm the reliability of the
experimental data.
The reasons for the small cooling capacity in Figure 5.11 may be made clear as the following
analysis results are understood.
300
200
m(kg/hr)
100
m_sol_GEN
m_sol_LPA
m_sol_MPA
0
12 16 20 24 28
T31(oC)
Figure 5.12 Solution flows vs. chilled water temperature
Figure 5.12 shows variation of the solution flow rates to the generator and both absorbers
during the tests. All three flow rates decrease as the chilled water temperature decrease. This
is because the control frequency for the solution pump and the flow control valve openings
for the absorbers were set constant during the tests and the working conditions for the system
changed in such a way as to decrease the flow rates. For example, system pressures changed
as in Figure 5.13.
10
7
p(kPa)
2.4
2
Phigh
1.6
Pmid
1.2 Plow
12 16 20 24 28
T31( C)
o
In Figure 5.13, the pressure differentials decrease as the chilled water temperature decreases
contributing to the decreasing flow rates in Figure 5.12.
Another reason for the decreasing flow rates can be found in the following concentration
curves.
In Figure 5.14, it is shown that all three concentrations increase with decreasing chilled water
temperature. Increase of concentration on the solution side, i.e. in the generator and
absorbers, means the refrigerant accumulates on the refrigerant side, which also means that
less volume of solution is in the solution lines. Consequently, due to the lack of solution to
form the hydraulic head, the flow rates of solution from the generator decrease.
0.58
0.56
0.54
0.52
0.5
xLiBr
0.48
GEN outlet
GEN inlet
0.16 refrigerant
0.08
0
12 16 20 24 28
T31(oC)
Figure 5.14 Concentrations vs. chilled water temperature
Besides the accumulation of refrigerant in the refrigerant tank, the increase of LiBr
concentration of the refrigerant in Figure 5.14 confirms that the content in the refrigerant
tank, i.e. the liquid level in the tank is increased as the chilled water temperature decreases.
Since, during the tests, the control valve openings and the frequency setting for the
refrigerant pump were also maintained constant, on the contrary to the decreased solution
flow rates, the refrigerant flow rates increases as the chilled water temperature decreases as
in the following Figure 5.15.
80
60
mref(kg/hr)
40
20
m_ref_total
m_ref_LPE
m_ref_MPE
0
12 16 20 24 28
T31(oC)
Figure 5.15 Refrigerant flow rates vs. chilled water temperature
The reason of all these changes in solution and refrigerant flow rates, i.e. the accumulation of
refrigerant on refrigerant side, is thought to originate from the decreasing cooling capacity
for decreasing chilled water temperature in Figure 5.11. Because the cooling capacity, in
other words, the absorption of refrigerant vapor to the absorber is limited by chilled water
temperature, until the generator and condenser begin to feel the influence, they keep
separating refrigerant from the solution at the same rate, which means more incoming
refrigerant flow than the evaporated refrigerant from evaporators. Therefore the refrigerant
tank should contain more and more refrigerant as the cooling capacity decreases. This
reasoning can be supported by comparison of the condensation and evaporation rates as in
Figure 5.16.
25
20
m(kg/hr)
15
10
5
m_cond
m_evap
0
12 16 20 24 28
T31(oC)
Figure 5.16 Condensation and evaporation rate vs. chilled water temperature
The condensation rate in Figure 5.16 was determined from the condensation heat calculated
from the measurement. The evaporation rate is the sum of the vapor flow rates in the two
evaporators determined from the heat transfer rates. As is clearly shown in Figure 5.16, the
condenser provides much more refrigerant flow rate than evaporated in the evaporators. The
difference between the condensation and evaporation rates returns to the solution side not via
‘evaporation/absorption’ process but for example, by simple mixing at the bottom of LPA as
illustrated in Figure 5.5. This is certainly a huge waste of energy seriously deteriorating
system COP as shown in Figure 5.17.
0.25
0.2
0.15
COP
0.1
0.05
0
12 16 20 24 28
T31(oC)
Figure 5.17 COP vs. chilled water temperature
COP in the Figure 5.17 is purely thermal and thus does not include auxiliary powers. COP
varies from 0.21 to 0.067 as the chilled water temperature decreases, significantly low
compared to the design COP of 0.38. The maximum COP of 0.21 is mainly due to the loss of
refrigerant. But the explanation of the even lower COP in the low chilled water temperature
region (also the sharp decrease of cooling capacity in Figure 5.11) needs more background
information as follows.
The first reason for the low COP and small capacity is the impurity in refrigerant. In Figure
5.14, the LiBr concentration in refrigerant increases from 4.5 to 7% as the chilled water
temperature decreases from 24.5 to 13°C. The tendency of LiBr concentration in refrigerant
is not so important and even may not have any regular pattern at all. More important is that
infusion of the solution into refrigerant is large enough to change not only the evaporation
temperature but also the transport properties of the refrigerant negatively.
The second but perhaps the biggest reason is the small recirculation flow rates for the
evaporators as explained in section 5.1.3. With small refrigerant flow rates, it is simply
impossible to use the whole evaporator surfaces effectively. This is especially true with
MPE, which has a vertical bare surface with a wide perimeter to be wetted with the
refrigerant.
From the analysis, it turned out that influence of the changes in evaporation temperature is
not so significant. During the whole tests, LiBr concentration in refrigerant varied from 0 to
16.5 %. The increase of equilibrium temperature during evaporation process, i.e. temperature
glide, was maximum 2.55 K and the average was 0.17 K for LPE. And for MPE, they were
2.88 and 0.22 K respectively.
The reason for the poor system performance can be best understood from the following
analysis of heat transfer coefficients.
10 0.6
8
hf_CON(kW/m2K)
hf_GEN(kW/m2K)
0.4
6
4
0.2
0 0
12 16 20 24 28 12 16 20 24 28
T31(oC) T31(oC)
(a) Condenser (b) Generator
Figure 5.18 Heat transfer coefficients of high-pressure components
In Figure 5.18, as the chilled water temperature decreases, heat transfer coefficients of the
condenser and generator also slowly decrease.
The decrease of condensation heat transfer coefficient is due to the decreasing condensation
rate, which retards the expose of the fresh condensation surface to vapor.
Decrease of the generator’s heat transfer coefficients is due to the decrease of the solution
flow as shown in Figure 5.12, which means increased dry surface area.
Neither of the decreasing heat transfer coefficients is thought to have caused the poor system
performance directly because the refrigerant flow rates from these components were always
larger than needed to yield the small cooling capacities as shown in Figure 5.16.
It should be mentioned that the heat transfer coefficients in Figure 5.18 were determined
directly from the measured bulk temperatures, not the saturation temperature calculated from
the bulk concentrations and the pressure. But this approach was not possible for the
evaporators and absorbers, where temperature profiles were observed to cross and mixing of
the refrigerant and solution take place. The attempts to determine the heat transfer
coefficients of LPE and MPA based on LMTDs failed completely.
Therefore, for the analysis of evaporators and absorbers, the measured bulk temperatures on
the solution and refrigerant side are not used but instead the saturation temperatures are used
in the calculation of LMTDs.
Figure 5.19 shows the heat transfer coefficients of low-pressure components, i.e. LPE and
LPA. The heat transfer coefficient of LPE increases as the chilled water temperature
decreases in 5.19a. Increasing refrigerant flow rate to LPE in Figure 5.15 may be the reason
for this but seeing that the gradient of the flow rate is very slow, it is hard to believe that it is
the main reason for such a change in the heat transfer coefficients as in Figure 5.19a. It is
unclear at this moment whether this trend is the results of measurement errors or not. No
matter what the reason is, the change in the performance of LPE must not have contributed to
the poor system performance.
2 1.6
1.6
hf_LPE(kW/m2K)
hf_LPA(kW/m2K)
1.2
1.2
0.8
0.8
0.4
0.4
0 0
12 16 20 24 28 12 16 20 24 28
T31(oC) T31(oC)
(a) LPE (b) LPA
Figure 5.19 Heat transfer coefficients of low-pressure components
On the other hand, the heat transfer coefficient of LPA decreases as the chilled water
temperature decreases. It originates from the fact that the solution flow rate was decreased as
shown in Figure 5.12. It should be mentioned that the heat transfer coefficient of LPA was
calculated by the empirical correlation Equation(4.30) determined from falling film tests in
Ch. 4. This was because one of the heat transfer coefficients of LPA and MPE should be
somehow determined firstly before the other and the LPA’s surface was observed completely
wet during the tests while the refrigerant flow rate in MPE was so small that even distribution
of the refrigerant in MPE could not be guaranteed. LPA’s heat transfer coefficients
determined from the MPE’s heat transfer coefficients, which was calculated by the empirical
correlations from the falling film tests, were too small to be accepted.
Seeing the order of magnitude, neither of the changes in LPE and LPA should be the primary
reason for such sharp deterioration of the system performance as in Figure 5.11 and 5.17
Performance changes of mid-pressure components, i.e. MPE and MPA, were observed
greater than the others. Figure 5.20 shows the heat transfer coefficients of MPE an MPA.
0.16
0.6
hf_MPE(kW/m2K)
hf_MPA(kW/m2K)
0.12
0.4
0.08
0.2
0.04
0 0
12 16 20 24 28 12 16 20 24 28
T31(oC) T31(oC)
(a) MPE (b) MPA
Figure 5.20 Heat transfer coefficients of mid-pressure components
Heat transfer coefficients of both components sharply decrease as chilled water temperature
decrease.
On the contrary to LPE in Figure 5.19a, the heat transfer coefficient of MPE in Figure 5.20
significantly decreases with decreasing chilled water temperature even when the refrigerant
flow rate increases as in 5.15. It is hard to explain why two contradictory trends exist in these
two evaporators. It may be due to the relatively large measurement errors for the small heat
transfer rates (QLPE=0.72 kW and QMPE=1.2 kW at T31=13°C) in the low chilled water
temperature region. Or it may be simply the difference between the two evaporators’ type
(LPE is a spiral coil and MPE is a vertical cylinder). Although the reason of the two different
trends is unclear at this moment, the small heat transfer coefficient of MPE (max. hf_MPE=132
W/m2K in Figure 5.20a) must have contributed to the poor system performance.
Similar to the trend in Figure 5.19a, the heat transfer coefficient of MPA also significantly
decreases with decreasing chilled water temperature. Seeing the change of the solution flow
rate to MPA in Figure 5.12, MPA seems to experience serious reduction in wetted area in the
flow range.
As a conclusion, the poor performance of both MPE and MPA seems to have caused the
decrease in the cooling capacity and the COP shown in Figure 5.11 and 5.17 respectively.
The primary reasons for the poor performance MPE and MPA seem to be the small flow
rates. Since the refrigerant flow rate to MPE cannot be increased due to the cavitation of the
irreplaceable refrigerant pump in Section 5.1.3, this point cannot be shown nor its
performance can be improved at this moment. MPA’s performance, however, can be
improved in its performance by increasing the solution flow rate. This will be shown in the
following sections.
In this section, the influence of chilled water temperature on the performance of the overall
system and the components was investigated. The reason why unusually high chilled water
temperature was used during the tests was explained. And finally, some of the components
were suspected to have caused the poor system performance.
The influences of heating medium and coolant temperature on the performance of the system
and the components are investigated.
Heating medium, i.e. hot water in this case, temperature is probably the most interesting
parameter of all because the goal of the present project is to develop an effective machine
working at the lowest heating medium temperature as possible.
Since the generator temperature is a very strong function of coolant (cooling water)
temperature, influences of both heating medium and coolant temperatures are presented
together in this section.
Figure 5.21 shows the influence of the hot water temperature on the cooling capacity at
different coolant temperatures.
3
QLPE(kW)
T33
1
40oC
37oC
33oC
0
50 60 70 80 90
T35(oC)
Figure 5.21 Cooling capacity vs. heating medium temperature
It should be noted that T35 in Figure 5.21 is the heating medium temperature at the generator
inlet. T33 is the lowest coolant temperature in the system measured at the inlet of MPA
because the coolant flows through MPA and condenser in series.
It seems that the cooling capacity approaches a certain limit value as the heating medium
temperature increases for all coolant temperatures in Figure 5.21. The gradients of the
capacity curves are steep in the low heating (medium) temperature region. At a constant
heating temperature, in low heating temperature region, the cooling capacity increases as the
coolant temperature decreases. The increase of cooling capacity is greater when T33 is
decreased from 37 to 33°C than that of 40 to 37°C.
In the high heating temperature region above 85°C, the cooling capacity seems, however, not
to be influenced by the coolant temperature.
For the coolant temperature of 37°C, the heating medium temperature should be around 80°C
to achieve 80% of the maximum capacity (QLPE=3.7 kW at 88.5°C).
Thermal system COPs were calculated and the results are presented in Figure 5.22. It should
be mentioned that the heat loss determined from the overall heat balance ranged from 0.7 to
2.8 kW in this series of tests.
0.3
COP
0.2
0.1 T33
40oC
37oC
33oC
0
60 70 80 90
T35( C)
o
It seems that there is a certain heating medium temperature that yields the best COP for each
coolant temperature. In this series of tests, COP ranged from 0.17 to 0.24 yielding an average
of 0.22. Unlike the cooling capacity, COP of this particular setup seems relatively insensitive
to the heating medium and coolant temperatures except for the T33=37°C case. The
decreasing gradient of COP below the heating temperature 75°C for 37°C coolant
temperature is exceptionally steep. This may be because of the exceptionally high LiBr
concentration in refrigerant at the low heating medium temperature as shown in Figure 5.23.
0.25
T33
40oC
0.2 37oC
33oC
0.15
xLiBr
0.1
0.05
0
60 70 80 90 100
T35(oC)
Figure 5.23 LiBr concentration in refrigerant vs. heating medium temperature
It seems that the exceptionally high concentration in refrigerant was unnoticed during the
test. Otherwise, exercise of a single blow down procedure would have lowered the
concentration to a comparable level of the other cases.
Figure 5.24 shows the variation of the system pressures during the tests.
For a coolant temperature, high pressure increases linearly with increasing heating
temperature in Figure 5.24a.
Mid- and low pressures are, however, relatively insensitive to the changes in heating medium
temperature. Since these pressures are complex functions of conditions on the solution side,
cooling capacity etc, the trends in Figure 5.24a and 5.24b should be analyzed together with
other parameters that follow.
12
T33 T33
T33
3.2 40oC 2.8 40oC
11 40oC
37oC 37oC
37oC
33oC 33oC
33oC
10
phigh(kPa)
pmid(kPa)
plow(kPa)
2.8
2.4
9
2.4
8 2
2
7
1.6 1.6
6
50 60 70 80 90 50 60 70 80 90 50 60 70 80 90
T35(oC) T35(oC) T35(oC)
(a) High pressure (b) Mid pressure (c) Low pressure
Figure 5.24 System pressures vs. heating medium temperature
Figure 5.25 shows the variation of working fluids’ flow rates during the tests. The refrigerant
flow rates from the condenser increase with increasing heating medium temperature in Figure
5.25a, which can be associated with the trends in Figure 5.24a.
24 180 200
20 160 180
m(kg/hr)
m(kg/hr)
m(kg/hr)
16 140 160
12 120 140
The solution flow rates, however, increase as the heating temperature decreases in Figure
5.25a and 5.25b. This is because the solution column, i.e. hydraulic head, builds up at
downstream of the generator as less refrigerant is generated due to the decreasing heating
medium temperature. But Figure 5.25b and 5.25c do not seem clearly associated with Figure
5.24b and 5.24c.
The trends of mid- and low pressures in Figure 5.24b and 5.24c seem more related to the
refrigerant flow rates to the evaporators as shown in Figure 5.26. In Figure 5.26, it can be
seen that the gradients of the flow rates resemble those of pressures in Figure 5.24. This fact
also suggests that the role of refrigerant recirculation in this particular setup is very
important.
50 40
40
30
m(kg/hr)
m(kg/hr)
30
20
20
T33 10 T33
10 40oC 40oC
37oC 37oC
33oC 33oC
0 0
60 70 80 90 60 70 80 90
T35(oC) T35(oC)
(a) MPE (b) LPE
Figure 5.26 Refrigerant flow rates vs. heating medium temperature
The refrigerant flow rates to both evaporators decrease sharply with decreasing heating
medium temperature. This is because the refrigerant is generated less as the heating
temperature gets lower and also the liquid level in the refrigerant tank is reduced so that
cavitation problem of the refrigerant pump should worsen.
Figure 5.27 shows the heat transfer coefficients of the condenser and generator. The heat
transfer coefficient curves in the figure resemble the corresponding flow rates in Figure 5.25.
12 0.6
hf_CON(kW/m2K)
hf_GEN(kW/m2K)
8 0.4
4 0.2
T33 T33
40oC 40oC
37oC 37oC
33oC 33oC
0 0
60 70 80 90 60 70 80 90
T35(oC) T35(oC)
(a) Condenser (b) Generator
Figure 5.27 Heat transfer coefficients of high-pressure components
Since there is no sharp change in the heat transfer coefficients of the generator in Figure
5.27b, any change in the generator’s performance is thought to have contributed to the steep
gradient of the cooling capacity curve in Figure 5.21. Neither the condenser’s heat transfer
coefficient should have caused it. Because the condenser’s heat transfer coefficient is not
independent parameter, i.e. dependent variable to the vapor generation in the generator, the
condenser cannot be blamed.
Following Figure 5.28 shows the heat transfer coefficients of LPE and LPA.
1 1.7
T33 T33
40oC 40oC
0.8 37oC 1.6 37oC
hf_LPE(kW/m2K)
hf_LPA(kW/m2K)
33oC 33oC
0.6 1.5
0.4 1.4
0.2 1.3
0 1.2
60 70 80 90 60 70 80 90
T35(oC) T35(oC)
(a) LPE (b) LPA
Figure 5.28 Heat transfer coefficients of low-pressure components
It can be seen that the heat transfer coefficients of the condenser and generator are strong
functions of the solution flow rates in Figure 5.26b and 5.25c respectively.
In Figure 5.28a, the heat transfer coefficients of LPE decrease roughly by 75% (from the
maximum 800 to 200 W/m2K), as the heating medium temperature decreases from 90 to 65
o
C. This is comparable to the reduction of refrigerant flow rate in Figure 5.26b. This strongly
suggests direct relationship between the flow rate and wetted surface area.
On the contrary, the heat transfer coefficients of LPA increase as the heating medium
temperature decreases. This is due to the increasing solution flow rate with decreasing
heating medium temperature as shown in Figure 5.25c.
Finally, Figure 5.29 shows the heat transfer coefficients of MPE and MPA.
0.25 1
0.2 0.8
hf_MPE(kW/m2K)
hf_MPA(kW/m2K)
0.15 0.6
0.1 0.4
T33 T33
0.05 40oC 0.2 40oC
37oC 37oC
33oC 33oC
0 0
60 70 80 90 60 70 80 90
T35( C)
o
T35( C)
o
Although magnitude of the changes is smaller than that of LPE, MPE’s heat transfer
coefficients decrease substantially as the heating medium temperature decreases. Depending
on the coolant temperature, it decreases by 30 to 50% as the heating medium temperature
decreases from 90 to 65°C as in Figure 5.29a.
Unlike the LPA in Figure 5.28b, MPA’s heat transfer coefficient shows different
characteristics even though the solution flow rates of the two absorbers follow a similar
pattern as shown in Figure 5.25b and 5.25c.
From the analysis above, the first and most critical reason of the small and rapidly decreasing
cooling capacity in Figure 5.21 is the heat transfer coefficients of the two evaporators, small
and steeply decreasing with the decreasing heating medium temperature.
Behind the poor performance of the evaporator are, again, the small recirculation flow rates
that are limited by the cavitation of the refrigerant pump as shown in Figure 5.26 and also, in
this particular series of tests, the sharply decreasing refrigerant generation as shown in Figure
5.25a. Compared to the influences of the refrigerant-side parameters, the other parameters’
influences are believed to be marginal.
This section describes the results acquired from a series of tests where the heating medium
flow rate was varied while the power input to the heater control was kept at 50% (nominal
power of 15 kW). This constant heat input was intended to simulate a real solar cooling
system, which has fixed area of solar collectors, because the capacity of a solar collector is
relatively insensitive to the change in heating medium flow rate within the test range. It
should be noted that the heating medium temperature is not a controllable parameter in this
case.
Figure 5.30 shows the variation of cooling capacity during the tests.
3.2 0.25
0.2
2.8
QLPE(kW)
0.15
COP
2.4
0.1
0.05
2
0
800 1200 1600 2000 2400 800 1200 1600 2000 2400
mheat.med.(kg/hr) mheat.med.(kg/hr)
(a) Cooling capacity (b) COP
Figure 5.30 Cooling capacity vs. heating medium flow rate
Both cooling capacity and COP approach certain values with decreasing heating medium
flow rate, which is good for the system economics. Although the power input was kept at 15
kW in the control program, the net heat gain by the generator was varied slightly as in Figure
5.31.
15.2
14.8
QGEN(kW)
14.4
14
13.6
As the heating medium flow rate decreased, overall heat transfer coefficient should have
decreased. This must have caused the net heat input to the generator to decrease in Figure
5.31. The refrigerant flow rate shows, however, quite different trend as in Figure 5.32.
20
19
m(kg/hr)
18
17
16
15
800 1200 1600 2000 2400
mheat.med.(kg/hr)
Figure 5.32 Condensation rate vs. heating medium flow rate
On the other hand, the solution and refrigerant recirculation flow rates were changed as in
Figure 5.33 during the tests.
80 400
60 300
m(kg/hr)
m(kg/hr)
40 200
20 100
m_ref_total m_sol_GEN
m_ref_LPE m_sol_LPA
m_ref_MPE m_sol_MPA
0 0
800 1200 1600 2000 2400 800 1200 1600 2000 2400
mheat.med.(kg/hr) mheat.med.(kg/hr)
(a) Refrigerant recirculation (b) Solution
Figure 5.33 Cooling capacity vs. heating medium flow rate
Refrigerant recirculation rates decrease with decreasing heating medium flow rate in Figure
5.33a. On the contrary, like the previous cases, the solution flow rates increase instead in
Figure 5.33b.
As can be expected, the heat transfer coefficients of the two absorbers slightly increase with
decreasing heating medium flow rate as in Figure 5.34.
1.4 0.8
1.35
hf_MPA(kW/m2K)
hf_LPA(kW/m2K)
0.6
1.3
1.25
0.4
1.2
1.15 0.2
800 1200 1600 2000 2400 800 1200 1600 2000 2400
mheat.med.(kg/hr) mheat.med.(kg/hr)
(a) LPA (b) MPA
Figure 5.34 Heat transfer coefficients of absorbers
But the heat transfer coefficients of the two evaporators did not decrease monotonously with
decreasing heating medium flow rate. In the following Figure 5.35, in spite of the
monotonously decreasing refrigerant flow rates in Figure 5.33a, LPE’s heat transfer
coefficient curve has an inflection point and MPE’s heat transfer coefficient even increases.
1.4 0.2
0.16
hf_MPE(kW/m2K)
hf_LPE(kW/m2K)
1.2
0.12
1
0.08
0.8
0.04
0.6 0
800 1200 1600 2000 2400 800 1200 1600 2000 2400
mheat.med.(kg/hr) mheat.med.(kg/hr)
(a) LPE (b) MPE
Figure 5.35 Heat transfer coefficients of evaporators
These trends may be due to the variation of the refrigerant concentration in Figure 5.36.
0.56
0.54
0.52
0.5
T33=37oC
xLiBr
0.48 rich solution
poor solution
0.06 refrigerant
0.03
0
800 1200 1600 2000 2400
mheat.med.(kg/hr)
Figure 5.36 LiBr concentrations vs. heating medium flow rate
No matter what the reason is, the increasing cooling capacity and COP with decreasing
heating medium flow rate in Figure 5.30 must have originated from the increased heat
transfer coefficients in Figure 5.34 and 5.35.
The influence of chilled water flow rate on the system performance is investigated in this
section.
3.2
0.25
2.8
QLPE(kW)
0.2
COP
2.4 0.15
0.1
2
0.05
800 1000 1200 1400 1600 800 1000 1200 1400 1600
Gchilled water (kg/hr) Gchilled water(kg/hr)
(a) Cooling capacity (b) COP
Figure 5.37 System performance vs. chilled water flow rate
Figure 5.37 shows the variation of the cooling capacity and COP while the chilled water flow
rate is changed. Both values decrease slightly as the flow rate is decreased.
In order to make sure that this is not the influence of LiBr concentration in the refrigerant, the
concentration is firstly checked in Figure 5.38.
0.56
0.54
0.52
0.5
xLiBr
0.48 T33=37oC
rich solution
poor solution
refrigerant
0.09
0.06
800 1000 1200 1400 1600
Gchilled water(kg/hr)
Figure 5.38 LiBr concentrations vs. chilled water flow rate
As is shown in the figure, since LiBr concentration increases with decreasing chilled water
flow rate, the results in the Figure 5.37 are also not free from the uncertainties regarding the
composition of refrigerant. But because the system performance is expected to deteriorate
with decreasing chilled water flow rate, the increasing LiBr concentration would only
accelerate the expected deterioration.
The other experimental results are not presented here because no special results were found
to report. Compared to the previous cases, no significant changes in flow rates for the
working fluids and the system pressures were observed. Only the heat transfer coefficients of
the two evaporators decreased slightly in the region of low chilled water flow rate where the
LiBr concentration is high.
The deteriorating system performance in Figure 5.37 is mainly due to the decreased overall
heat transfer coefficient of LPE as the chilled water flow rate is decreased.
Since the chilled water temperature is more or less determined by the cooling requirement
and its influence on the system performance is not strong as shown above, the chilled water
flow rate would better be determined by the cooling capacity and the required chilled water
temperature.
The influence of coolant (cooling water) flow rate on the system performance is investigated
in this section.
Figure 5.39 shows the variation of the cooling capacity and COP while the coolant flow rate
is changed. Both values decrease slightly as the flow rate is decreased.
5 0.4
4
0.3
QLPE(kW)
3
COP
0.2
2
0.1
1
0 0
1200 1600 2000 2400 1200 1600 2000 2400
mcoolant(kg/hr) mcoolant(kg/hr)
(a) Cooling capacity (b) COP
Figure 5.39 System performance vs. coolant flow rate
It turned out that the system is rather insensitive to the coolant flow rate within the tested
range.
Since the LiBr concentration in refrigerant was maintained at more or less a constant value as
in Figure 5.40, the trends in Figure 5.39 are not sure in relation to the impurity of refrigerant.
0.56
0.54
0.52
0.5
xLiBr
0.48 T33=37oC
rich solution
poor solution
0.1 refrigerant
0.08
Figure 5.41 shows variations of the solution flow rates and the system pressures during the
tests.
All solution flow rates increase with decreasing coolant flow rate in Figure 5.41a. This is
partly because the increasing pressure differences between the system high pressure and the
other system pressures in Figure 5.41b.
400 10.8
10.2
300 9.6
m(kg/hr) 9 T33=37oC
p(kPa)
8.4 Phigh
200 Pmid
Plow
2.4
100
m_sol_GEN
2.2
m_sol_LPA
m_sol_MPA
0 2
1200 1600 2000 2400 1200 1600 2000 2400
mcoolant(oC) mcoolant(kg/hr)
(a) Solution flow rates (b) Pressures
Figure 5.41 System performance vs. coolant flow rate
It is worth mentioning that the system mid pressure is lower than the system low pressure for
most of the test range. This is because the chilled water temperature is kept at the rather high
temperature as explained in section 5.2.1 in order to maintain reasonably large heat transfer
rates in all components. As the coolant temperature is decreased, the high- and mid pressures
increase and the mid pressure becomes equal to the low pressure at the coolant flow rate of
1200 kg/hr.
Figure 5.42 shows the variation of refrigerant flow rates during the tests. As the system high
pressure increases with decreasing coolant flow rate, the refrigerant generation decreases as
in Figure 5.42a. Consequently all the refrigerant recirculation rates also decrease in Figure
5.42b.
20 80
16
60
m(kg/hr)
m(kg/hr)
12
40
8
20
4 m_ref_total
m_ref_LPE
m_ref_MPE
0 0
1200 1600 2000 2400 1200 1600 2000 2400
mcoolant(kg/hr) mcoolant(kg/hr)
(a) Condensation rates (b) Refrigerant recirculation rates
Figure 5.42 Refrigerant flow rates vs. coolant flow rate
But it seems that the reduction of refrigerant generation is not large enough to significantly
change the heat transfer coefficients of the two evaporators in Figure 5.43.
1 0.2
0.8 0.16
hf_MPE(kW/m2K)
hf_LPE(kW/m2K)
0.6 0.12
0.4 0.08
0.2 0.04
0 0
1200 1600 2000 2400 1200 1600 2000 2400
mcoolant(kg/hr) mcoolant(kg/hr)
(a) LPE (b) MPE
Figure 5.43 Heat transfer coefficients of evaporators
On the other hand, the increasing solution flow rates in Figure 5.41a seem to have more
influence on the two absorbers as in Figure 5.44.
1.5 2
1.6
hf_MPA(kW/m2K)
hf_LPA(kW/m2K)
1.4
1.2
0.8
1.3
0.4
1.2 0
1200 1600 2000 2400 1200 1600 2000 2400
mcoolant(kg/hr) mcoolant(kg/hr)
(a) LPA (b) MPA
Figure 5.44 Heat transfer coefficients of absorbers
As a conclusion, these increasing heat transfer coefficients of the absorbers seem to have
compensated the deterioration of the system performance as the coolant flow rate was
decreased.
It is recommended that the minimum flow rate should be used to save pumping power for the
coolant pump as long as the system performance does not deteriorate substantially.
Because the system has two evaporators to deliver liquefied refrigerant, the distribution ratio
of the refrigerant is very interesting parameter. The influence of this parameter on the system
performance is investigated in this section.
The influences on the system performance can be observed in Figure 5.45. The results were
shown against the recirculation flow rate to LPE (Since all the solution flow rates and
refrigeration flow rates changed during the tests, ‘distribution ratio’ was not thought to be an
independent parameter).
While an optimum flow rate seems to exist for the cooling capacity in Figure 5.45a, COP
keeps increasing as less refrigerant is supplied to LPE within the test range as shown in
Figure 5.45b.
4 0.25
0.2
3
QLPE(kW)
0.15
COP
2
0.1
1
0.05
0 0
10 15 20 25 30 35 40 10 15 20 25 30 35 40
mref_LPE(kg/hr) mref_LPE(kg/hr)
(a) Cooling capacity (b) COP
Figure 5.45 System performance vs. refrigerant distribution
Figure 5.46a shows the all three refrigerant flow rates in the recirculation circuit, including
the total recirculation flow rate, the refrigerant flow rate to MPE and that of LPE. The total
recirculation flow rate increases as less refrigerant goes to LPE.
100 25
m_ref_total
m_ref_LPE
80 m_ref_MPE 20
m(kg/hr)
m(kg/hr)
60 15
40 10
20 5
m_cond
m_evap
0 0
10 15 20 25 30 35 40 10 15 20 25 30 35 40
mref_LPE(kg/hr) mref_LPE(kg/hr)
(a) Refrigerant recirculation (b) Condensation and evaporation rates
Figure 5.46 Refrigerant flow rates vs. refrigerant distribution
Figure 5.46b compares the liquefied refrigerant supplied from the condenser and the amount
of the refrigerant evaporated from the two evaporators. The difference of the two curves is
the refrigerant wasted and it can be seen that the difference is the minimum at the smallest
refrigerant flow rate to LPE. Physically, this is due to refrigerant loss as explained in section
5.1.3. This may explain the trend of COP in Figure 5.45b.
Figure 5.47 shows the corresponding variation of the two evaporators’ heat transfer
coefficients.
LPE’s heat transfer coefficient decreases as the refrigerant flow rate decreases. MPE’s heat
transfer coefficient seems, however, not to be influenced by the large change in its refrigerant
flow rate in Figure 5.47b.
1
0.8 0.12
hf_MPE(kW/m2K)
hf_LPE(kW/m2K)
0.6
0.08
0.4
0.04
0.2
0 0
10 15 20 25 30 35 40 10 15 20 25 30 35 40
mref_LPE(kg/hr) mref_LPE(kg/hr)
(a) LPE (b) MPE
Figure 5.47 Heat transfer coefficients of evaporators
Another point that should be noticed in Figure 5.47 is the difference in the absolute values of
two heat transfer coefficients. LPE’s heat transfer coefficient is 6 to 8 times larger than that
of MPE in spite of the relatively smaller refrigerant is supplied to LPE as shown in Figure
5.47a. This huge difference cannot be explained by possible errors in measurement or
analysis process because it is simply too large for such reasons. It may rather be due to the
difference in types of the evaporators.
In a coiled tube type evaporator like LPE, small drops of refrigerant have better chance to
flow a long way along the spiral path on the coiled tube. And splashing onto the surface of
the next turn of the tube also increases the heat transfer coefficient.
The vertical cylinder type evaporator like MPE has, however, no such mechanism in the
evaporation process unless the surface has been treated or processed specially for heat
transfer enhancement. This difference may cause the huge difference between the
performances of two evaporators.
Besides, in the present chiller setup, the refrigerant is contaminated by LiBr solution that
contains not only the LiBr salt but also octanol as a surface active agent. It will be shown in
Chapter 6 that the presence of octanol in pure water dramatically deteriorated the heat
transfer performance in single-phase heat transfer. Although evaporation is not a single-phase
heat transfer process, considering that the evaporation takes place only at the interface, the
octanol may also similarly influence the bulk film heat transfer coefficient in evaporation
process. This could also contribute to the difference.
Before presenting the absorbers’ heat transfer coefficients, the variations of solution flow
rates are shown in Figure 5.48.
400
m_sol_GEN
m_sol_LPA
m_sol_MPA
300
m(kg/hr)
200
100
0
10 15 20 25 30 35 40
mref_LPE(kg/hr)
Figure 5.48 Solution flow rates vs. refrigerant distribution
All three solution flow rates decrease as less refrigerant is recirculated to LPE in Figure 5.48.
This is mainly because the refrigerant accumulates in the refrigerant tank and thus the
hydraulic head at the downstream of the generator gets smaller. The accumulation of
refrigerant is supported by the increasing total refrigerant flow rate in Figure 5.46a.
1.4 0.6
1.35
hf_MPA(kW/m2K)
hf_LPA(kW/m2K)
0.4
1.3
1.25
0.2
1.2
1.15 0
10 15 20 25 30 35 40 10 15 20 25 30 35 40
mref_LPE(kg/hr) mref_LPE(kg/hr)
(a) LPA (b) MPA
Figure 5.49 Heat transfer coefficients of absorbers
Both heat transfer coefficients decrease slightly with decreasing refrigerant flow rate to LPE.
Distribution of the solution flow from the generator to two absorbers is important parameter.
The results are shown against one of the solution flow rates, not the distribution ratio.
5 0.4
0.3
QLPE(kW) 4
COP
0.2
3
0.1
2 0
130 140 150 160 170 180 190 130 140 150 160 170 180 190
msol_LPA(kg/hr) msol_LPA(kg/hr)
(a) Cooling capacity (b) COP
Figure 5.50 System performance vs. solution distribution
In Figure 5.50, it seems that both cooling capacity and COP reach the maximum values at the
same solution flow rate to LPA, 157 kg/hr. But decreasing gradient of the cooling capacity is
slower than that of COP below that flow rate.
Figure 5.51 shows the corresponding changes in the solution flow rates during the tests.
400
300
m(kg/hr)
200
100
m_sol_GEN
m_sol_LPA
m_sol_MPA
0
130 140 150 160 170 180 190
msol_LPA(kg/hr)
Figure 5.51 Solution flow rates vs. solution distribution
The solution flow rate supplied to the generator has its minimum value at around the same
solution flow rate of 157 kg/hr.
The results above suggest that the even distribution of solution provides the best system
performance. Figure 5.52 and Figure 5.53 explain the trend of COP in Figure 5.50b.
25
20
m(kg/hr)
15
10
5 m_ref_cond
m_ref_evap
m_ref_loss
0
130 140 150 160 170 180 190
msol_LPA(kg/hr)
Figure 5.52 Condensation and evaporation rates vs. solution distribution
Figure 5.52 shows refrigerant flow rate from the condenser, the amount of refrigerant
evaporated from the two evaporators and the difference between the two. It can be seen that
more refrigerant is supplied from the condenser as less solution goes to LPA. More
importantly, the difference between the condensation and the evaporation rates shows a
minimum around the solution flow rate of 160 kg/hr.
Figure 5.53 shows the corresponding heat input to the generator. It is the minimum at the
solution flow rate of 157 kg/hr.
20
18
QGEN(kW)
16
14
12
10
130 140 150 160 170 180 190
msol_LPA(kg/hr)
Figure 5.53 Heat input to generator vs. solution distribution
In the previous sections, the control valves in the solution lines were set at the constant
position, which means the flow rates constantly changed depending on the working
conditions during the tests. Then it is interesting to investigate a constant flow case to see
how the control of solution flow rates influence the system performance.
In this section, the test results attained from a series of tests where the control valves were
used to maintain all solution flow rates at constant values are presented.
In this series of tests, the solution flow rates to LPA and the generator are maintained at 160
and 324 kg/hr respectively while the heating medium temperature is varied with all the other
parameters maintained at the standard values.
Since the refrigerant flow rate is a dependent variable, it changes according to the system
performance. Therefore the last solution flow rate, the flow rate to MPA is also not
controlled and it is subject to change during the tests. These solution flow rates are shown in
Figure 5.54.
25 400
m_ref_cond
m_ref_evap
20 m_ref_loss
300
m(kg/hr)
m(kg/hr)
15
200
10
100
5 m_sol_GEN
m_sol_LPA
m_sol_MPA
0 0
65 70 75 80 85 90 65 70 75 80 85 90
T35(oC) T35(oC)
(a) Condensation and evaporation rates (b) Solution flow rates
Figure 5.54 Flow rates vs. heating medium temperature
In Figure 5.54a, condensation rate decreases as the heating medium temperature is decreased.
Consequently, the flow rate to MPA is increased while the flow rates to LPA and the
generator are kept constant as in Figure 5.54b.
Figure 5.55 shows variation of the corresponding cooling capacity and the COP.
In Figure 5.55a, the cooling capacity is increased slightly compared to the uncontrolled case
in high heating temperature region. This increase in cooling capacity is, however, believed
mainly due to the absolute difference in the solution flow rates.
More interestingly, in Figure 5.55b, inflection point of COP seems to be shifted to lower
heating temperature region, which is beneficial for the system economics. Magnitude of the
shift is about 7~8 K.
4 0.5
T33=37oC T33=37oC
controlled flows controlled flows
uncontrolled flows 0.4 uncontrolled flows
3
QLPE(kW)
0.3
COP
2
0.2
1
0.1
0 0
50 60 70 80 90 60 70 80 90
T35(oC) T35(oC)
(a) Cooling capacity (b) COP
Figure 5.55 Influence of solution flow rate control on the system performance
Although the benefit may seem to be too small for possibly expensive control measures,
since the present setup is far from the original design due to many practical problems, it is
risky to make final conclusions.
After all experimental cases in Table 5.1 had been carried out, 2 kg of extra 50 % solution
was charged into the system. In addition, 3.83 ml of octanol was added to make its averaged
concentration in the solution to be 200 ppm.
The reason of charging extra 2 kg of solution is to increase the circulation rate of the solution
at high heating temperatures, because the maximum solution flow rate to the generator has
been less than 300 kg/hr at around 90°C during the previous tests.
More charge of octanol was intended to observe any influence of its increased concentration
on the system performance.
Since the influences of the secondary fluids were already investigated in the previous
sections, only the important temperature parameters are investigated in this section.
The following sets of tests in Table 5.2 have been carried out.
Table 5.2 Parameters and the test values (30 kg of 50% solution + 200 ppm octanol)
Parameters Step 1 Step 2 Step 3 Step 4 Step 5 Step 6 Remark
T31 (°C) 13 16 19 24 chilled water temp.
T35 (°C) 65 70 75 79 84 89 heating medium temp.
T33 (°C) 28 34 37 42 coolant temp.
1. Standard values are in the gray cells, to which the other variables are set while a variable is changed.
2. Numerical indices should be referred to Figure 5.1
3. Step vales other than standard ones are nominal. The real measured value can be different.
4. The solution from the generator is distributed to two absorbers as evenly as possible.
Figure 5.56 shows the variation of the system performance while the chilled water
temperature is varied. Compared to the results of the 1st charging condition, both cooling
capacity and COP are significantly improved.
5 0.25
4 0.2
QLPE(kW)
3 0.15
COP
2 0.1
The extra charging of the solution significantly widened the range of solution flow rates.
Maximum solution flow rates to absorbers were far beyond the design flow rates and the
solution flow rates during this set of tests are much steadier as in Figure 5.57 than those of
the 1st charging condition in Figure 5.12.
400
300
m(kg/hr)
200
100
m_sol_GEN
m_sol_LPA
m_sol_MPA
0
12 16 20 24
T31( C)
o
The solution flow rates decrease slightly with decreasing chilled water temperature in Figure
5.57.
Figure 5.58 compares the condensation rates for the cases of 1st and 2nd charging conditions.
The refrigerant generation of the 2nd charging case has been increased by 16~35% from the
1st case.
80 25
20
60
m(kg/hr)
m(kg/hr)
15
40
10
20
m_ref_total 5 T33=37oC
m_ref_LPE
2nd charging
m_ref_MPE
1st charging
0 0
12 16 20 24 12 16 20 24 28
T31(oC) T31(oC)
(a) Refrigerant recirculation (b) Condensation rate
Figure 5.58 Refrigerant flow rates vs. chilled water temperature
This is mainly because of the larger solution flow rates in Figure 5.57 than those of Figure
5.12. Increased solution flow rate means larger thermal capacities and higher transfer
coefficients for components.
Although the addition of extra octanol also must have played a role, the contribution must be
much less than the increased flow rate, which is supported by Figure 5.59. Figure 5.59 shows
variation of the Nusselt number of MPA from 58 experimental data against film Reynolds
number for different octanol concentrations.
0.24
Nuf
0.2
0.16
Octanol concentration
200ppm
100ppm
32 36 40 44 48
Ref
Figure 5.59 Influence of octanol concentration on MPA’s Nusselt number
Nusselt number for 200 ppm seems a little larger than that of 100 ppm in low Reynolds
number region. But the difference between the two sets of data is thought too small to bring
any practical difference.
Although the influence of increased octanol on LPA, however, may be greater, an attempt to
evaluate the influence in terms of Nusselt number failed. It is because there is no reliable
information for estimating the LPA’s Nusselt number for 200 ppm cases. Use of the
empirical correlations from the falling film tests for MPE [Equation (4.30)] gave
unacceptably small (even negative) Nusselt number for LPA, which was also the case in
Section 5.2.1.
comparing the subcooling (Tsat – Tbulk) of solutions at the bottom of two absorbers because
the subcooling is a measure of heat transfer coefficient. Figure 5.60 shows the subcooling in
the absorbers.
16 25
20
12
Ts21-T21(K)
Ts18-T18(K)
15
8
10
4
Octanol concentration 5 Octanol concentration
200ppm 200ppm
100ppm 100ppm
0 0
32 36 40 44 48 20 30 40 50
Ref Ref
(a) MPA (b) LPA
Figure 5.60 Subcooling of solution at the absorber outlet
For both absorbers, increased concentration of octanol decreases the subcooling of solution at
the outlet. The extra subcooling decreases with increasing Reynolds number for MPA but it
is constant for LPA. The magnitude of extra subcooling ranged from 0.6 to 1.5 K in MPA
and it is about 2 K in LPA. Therefore it may be said that the influence of increased octanol is
greater. But considering the increase of Nusselt number in MPA in Figure 6.59 is less than
5%, the improvement on LPA is comparable and thus can be neglected for purpose of the
present analysis.
The results of heat transfer analysis are as follows. Compared to the counterparts of the 1st
charging cases, heat transfer coefficient has been increased for all components.
Figure 5.61 shows the heat transfer coefficients of the generator and condenser.
T33=37oC
2nd charging 10
1st charging
hf_CON(kW/m2K)
hf_GEN(kW/m2K)
0.8
0.4
6
T33=37oC
2nd charging
1st charging
0 4
12 16 20 24 28 12 16 20 24
T31(oC) T31(oC)
(a) Generator (b) Condenser
Figure 5.61 Heat transfer coefficients of high-pressure components
In Figure 5.61a, the heat transfer coefficient of the generator follows the trends of the
corresponding solution flow rates in Figure 5.12 and 5.57.
hf_MPE(kW/m2K)
hf_LPE(kW/m2K)
1.2
0.2
0.8
0.1
0.4
T33=37oC
2nd charging
1st charging
0 0
12 16 20 24 28 12 16 20 24 28
T31(oC) T31(oC)
(a) LPE (b) MPE
Figure 5.62 Heat transfer coefficients of evaporators
Strange are the different trends between the two curves in Figure 5.62a. For both cases, the
refrigerant flow rate to LPE increased with decreasing chilled water temperature as in Figure
5.15 and 5.58a. Under the 1st charging condition, the heat transfer coefficient increased
sharply with decreasing chilled water temperature below 20°C but it decreased monotonously
with the 2nd charging condition. Considering the slow gradients of refrigerant flow rates in
Figure 5.15 and 5.58a, the heat transfer coefficient curve of the 1st charging condition is
questionable.
All the heat transfer coefficients in Figure 5.62b were calculated back from the LPA’s heat
transfer coefficients that have been calculated using the empirical correlation for the solution
with 100 ppm octanol from the falling film test.
Figure 5.63 compares the heat transfer coefficients of absorbers for different charging
conditions.
1.4
T33=37oC
2nd charging
1st charging
hf_MPA(kW/m2K)
hf_LPA(kW/m2K)
1.3 0.8
1.2
0.4
1.1
T33=37oC
2nd charging
1st charging
1 0
12 16 20 24 28 12 16 20 24 28
T31(oC) T31(oC)
(a) LPA (b) MPA
Figure 5.63 Heat transfer coefficients of absorbers
the setup attained larger capacity and higher transfer coefficients in the components, which
significantly improved the system performance.
In this section, the influence of heating medium temperature is investigated again on the
chiller setup under 2nd charging condition.
Figure 5.64 shows cooling capacity and the corresponding COP against heating medium
temperature.
5 0.4
T33=37oC
2nd charging
4 1st charging
0.3
QLPE(kW)
COP
0.2
2
0.1
1 T33=37oC
2nd charging
1st charging
0 0
50 60 70 80 90 60 70 80 90
T35(oC) T35(oC)
(a) Cooling capacity (b) COP
Figure 5.64 System performance vs. heating medium temperature
The cooling capacity increases linearly with increasing heating temperature for the 2nd
charging case in Figure 5.64a. The trend is quite different from the 1st charging case in high
temperature region, where the gradient becomes increasingly smaller.
The trends of COPs in Figure 5.64b are also different for the two cases. In the low
temperature region, while COP of the 1st charging case increases with increasing heating
temperature, that of 2nd charging decreases.
The cooling capacity has been increased by 24% at the design heating medium temperature
88.5°C (From 3.52 to 4.37 kW). COP is 0.255 at the design point, which is 115% of 0.222
from the 1st charging condition (increase of 0.033).
During this series of test, solution flow rate to the generator changed as in Figure 5.65. Under
the 1st charging condition, the flow rate increased sharply and then approached a constant
value as the heating medium temperature decreased. On the other hand, the flow rate
decreased linearly against heating temperature under the 2nd charging condition. The control
valves that regulate the solution flows to the two absorbers were set at the constant positions
during the tests for both cases.
400
360
m(kg/hr)
320
280
240
2nd charging
1st charging
200
70 80 90
T35(oC)
Figure 5.65 Solution flow rate to generator
The main reason of this performance improvement is mainly due to the increased heat
transfer coefficients of the components shown in Figure 5.66.
0.8 0.4
1.2
2nd charging 2nd charging
1st charging 1st charging
hf_MPE(kW/m2K)
1
hf_GEN(kW/m2K)
hf_MPA(kW/m2K)
0.3
0.6
0.8
0.2
0.6
0.4
0.1
0.4
2nd charging
1st charging
0.2 0 0.2
60 70 80 90 60 70 80 90 60 70 80 90
T35(oC) T35(oC) T35(oC)
(a) Generator (b) MPE (c) MPA
Figure 5.66 Improved heat transfer coefficients
It can be seen that, in Figure 5.66, increase of heat transfer coefficients of those components
is greater in high temperature region.
This section describes the influence of coolant temperature on the system working at several
different heating medium temperatures. The tests were carried out changing the coolant
temperature while heating medium temperature was maintained. The results are as follows.
Figure 5.67 shows variation of the cooling capacity and COP against the coolant temperature
for different heating medium temperatures.
6 0.4
QLPE(kW) 4 0.3
COP
2 0.2
T35 T35
88.5oC 88.5oC
79.5oC 79.5oC
70.5oC 70.5oC
0 0.1
28 32 36 40 44 48 28 32 36 40 44
T33(oC) T33(oC)
(a) Cooling capacity (b) COP
Figure 5.67 System performance vs. heating medium temperature
Both cooling capacity and COP decrease with increasing coolant temperature. The
decreasing gradient of cooling capacity is much steeper than that of COP.
The maximum cooling capacity of 5.3 kW was attained at the coolant and heating medium
temperature 31 and 88.5°C, respectively. The cooling capacity seems to increase linearly
with decreasing coolant temperature within the tested range.
COP ranged from 0.24 to 0.28 during this series of tests. Due to the lack of data, COP curve
for a certain heating medium temperature is hardly discernable from the others in Figure
5.66. But, for all cases, COP seems to increase linearly with decreasing coolant temperature.
Figure 5.68 shows the variation of the flow rates in the generator and condenser. Although
the solution flow rates in Figure 5.68a decrease with decreasing coolant temperature, the
condensation rates increase in Figure 5.68b.
360 28
24
340
20
m(kg/hr)
m(kg/hr)
320
16
300
12
T35 T35
280 88.5oC 88.5oC
8
79.5oC 79.5oC
70.5oC 70.5oC
260 4
28 32 36 40 44 48 28 32 36 40 44 48
T33(oC) T33(oC)
(a) Solution flow rate in generator (b) Condensation rate
Figure 5.68 Flow rates in high-pressure components
Figure 5.69 shows the corresponding heat transfer coefficients of the generator and
condenser.
0.8 12
hf_CON(kW/m2K)
hf_GEN(kW/m2K)
0.6
8
0.4
4
0.2 T35 T35
88.5oC 88.5oC
79.5oC 79.5oC
70.5oC 70.5oC
0 0
28 32 36 40 44 48 28 32 36 40 44 48
T33(oC) T33(oC)
(a) Generator (b) Condenser
Figure 5.69 Heat transfer coefficients of high-pressure components
The heat transfer coefficient of the generator approaches a maximum value as the coolant
temperature decreases for all cases in Figure 5.69a.
The condensation heat transfer coefficient decreases as the coolant temperature decreases for
all cases in Figure 5.69b.
It can be seen again how closely heat transfer coefficient of an evaporator is related to its
refrigerant recirculation in the following.
Figure 5.70a shows the refrigerant flow rate supplied to LPE and Figure 5.70b shows the
corresponding heat transfer coefficient. Although there are other parameters influencing the
heat transfer coefficient, one can see the strong resemblances between the two figures.
40 0.8
T35
88.5oC
79.5oC
hf_LPE(kW/m2K)
30 70.5oC 0.6
m(kg/hr)
20 0.4
0.2
T35
10
88.5oC
79.5oC
70.5oC
0 0
28 32 36 40 44 48 28 32 36 40 44 48
T33( C)
o
T33( C)
o
50 0.4
T35 T35
88.5oC 88.5oC
40 79.5oC 79.5oC
hf_MPE(kW/m2K)
70.5oC 70.5oC
m(kg/hr) 0.3
30
20 0.2
10
0.1
0
28 32 36 40 44 48 28 32 36 40 44 48
T33(oC) T33(oC)
(a) Solution flow rate to MPE (b) heat transfer coefficient
Figure 5.71 Refrigerant flow rate and heat transfer coefficient of MPE
Figure 5.70 and 5.71 reveal strong relationship between the performance of an evaporator
and its refrigerant flow rate.
Regarding the magnitudes of the heat transfer coefficient, that of MPE is unacceptably small
ranging from 100 to 300 W/m2K as shown in Figure 5.71b.
Although performance of the chiller setup has been significantly improved under the 2nd
charging condition, the cooling capacity and COP are still much lower than the original
target. the analysis indicates the poor performance of the evaporators as the reason for this
difference.
5.4 Conclusions
The half-effect chiller setup has been tested varying various parameters including
controllable temperatures, flow rates and charging condition. From the preliminary and main
test results, the following conclusions were made.
Regarding the erroneous designs and poor fabrication of the setup, the following problems
remain;
1. Air is constantly leaking into the system through the flange that connects the
condenser and the generator. The setup must be evacuated regularly as often as
possible.
2. Cavitation of the refrigerant gear pump significantly limits the circulation flow rates.
3. Leaking connections for solution and refrigerant flows contaminate and waste the
refrigerant deteriorating the system performance.
From the analysis of test results, the following conclusions were made;
4. By optimizing the charging amount and composition of LiBr solution, the system
performance could be improved significantly.
5. The present cooling capacity and COP are 4.36 kW and 0.25 meeting 44 and 66% of
the original goals (10 kW and 0.38 respectively). To this poor performance, the
following factors mainly contribute;
The reduction of the capacities of MPA and LPE are similar to the values already
expected in the final design (see 2nd progress report). The capacities of the generator
and LPA/MPE unit are particularly smaller than the design values.
A half-effect H2O/LiBr chiller has been successfully tested and the test results were fully
analyzed. Although the system performance is not close to what had been expected due to the
several unexpected and still unsolved problems, reasons for all those problems were
identified and valuable information has been acquired.
5.5 Recommendations
For any future research related to the present work, the following is recommended.
1. For the LPA/MPE unit, measures should be taken to improve the evaporator-side
heat transfer coefficient. This can be done either by recirculating enough refrigerant
or by using a special heat transfer surface on that side.
2. The generator should be redesigned or modified to increase either its film-side heat
transfer coefficient or to accommodate more heat transfer area. Also it is
recommended to reduce the height of condenser-generator column.
3. Design parameters for the tube coil heat exchangers, i.e. LPE and MPA, should be
reconsidered to avoid loss of refrigerant or solution, which include distribution, tube
pitch and the method of coil alignment.
4. Trays collecting solution or refrigerant should be positioned close to the last turn of
the tube coil and would better be made as big as possible.
If the refrigerant recirculation is to be used again, the circuit should be redesigned so that the
recirculation flow rate can be controlled freely and should not be influenced by the system
performance.
6.1 Start-up
Circulation of working fluids in the chiller setup heavily depends on the hydraulic heads built
up within the tubes between the components. Although spatial arrangement of those
components had been made carefully to guarantee design flow rates for the working fluids,
transient characteristic of this flow system was not completely understood until the
preliminary tests. From the troubleshooting during the preliminary tests, a procedure was
developed for stable start-up operation. Following is the description of the start-up procedure.
1
V3 V4
H1
from
evaporators
from
condenser
3 Solution
tank
V2 2
4
H4 H2
V1
HT-Hex
Refrigerant
tank
LT-Hex
Figure 6.1 Flow network in the setup
Figure 6.1 shows a schematic diagram of flow network and rough indication of the working
fluid levels in several components.
1. Heat up the heating medium to the desired testing temperature. (For a real machine, this is
equivalent to running a hot water pump.)
2. Open the valves, V1 and V2 to let the solution flow from the generator. Wait for some time
(30-40 seconds) to observe the solution flow at the bottom of LPA (point ③ in the figure)
through sight glasses. (From the previous shut-down process, the valves between the
generator and two absorbers are kept closed.)
3. Turn on the solution pump. Check the sight glass if there are two incoming solution flows
to the solution tank. (No incoming flow means either buildup of inert gas inside the solution
tank or insufficient solution charge.)
4. Run the refrigerant pump, chilled water- and cooling water pump. The order of operating
these components is not important.
5. Wait for the system to reach a stable condition and manipulate the valves, V1 ~V4 for
desired flow rates.
All the operations above can be programmed for a real machine according to an optimized
control scheme.
Solution flow rates to the generator and absorbers are most important parameters in
determining the capacity and COP of the setup. Without the valves, V1 and V2 in Figure 6.1,
those flow rates would have not been controllable parameters. Once a certain amount of
solution was charged into the setup, the flow rate from the generator is determined by the
sum of the pressure differentials and the hydraulic head H1 at the point ① in Figure 6.1.
Depending on the solution charge, the flow rate from the generator shows quite different
trends as shown in Figure 5.65. It seemed that the 1st solution charge was not enough to keep
a positive hydraulic head H1 in figure 6.1.
Figure 6.2 shows the solution flow rates from the generator for an uncontrolled (constant
openings for the valve V1 and V2 in Figure 6.1) and a controlled case (constant flow) under
the 1st charging condition. The changes in the system performance were already described in
section 5.2.8.
400
360
m(kg/hr)
320
280
Tcool=37oC
Controlled flow
240 Uncontrolled flow
60 70 80 90
T35( C)
o
Figure 6.2 Controlled and uncontrolled flows from the generator (1st charging condition)
As could be observed in Figure 5.55, capacity is slightly larger and COP shifts to lower
heating medium temperature with the constant flow case. But this does not necessarily
suggest the flow rate should be kept constant. It may be even better to reduce the flow rate
according to an optimized function as the heating medium temperature goes down.
For the control of the flow rate from the generator and its optimal distribution to two
absorbers, electric valves may be used. For example, wide range of solenoid-driven metering
valves is available in the market. The technology has long been proved for its diverse
applicability and reliability. It would not be difficult to integrate one into the setup.
Since the setup has many unintended features (e.g. contaminated refrigerant, loss of
refrigerant and etc.), any conclusion regarding the control of the solution flow rate may be
premature at this moment. Nevertheless, its influence is doubtlessly strong on the system
performance and thus its control should be considered with enough care.
Even though the operating points of the chiller are far from the crystallization limit of LiBr,
monitoring of the working condition is necessary for the potential risk. Actually it happened
once during the preliminary test in the high-temperature heat exchanger located downstream
of the generator, which is a typical place for crystallization. Although it happened in an
unusual working condition, importance of crystallization monitoring was reminded.
6.4 Shutdown
Since the highest operating concentration in the chiller is less than crystallization limit by
more than 10 LiBr wt%, danger of crystallization is relatively low compared to the other
LiBr systems. A proper shutdown procedure is important nevertheless.
1. Turn off the heating medium heater but keep the heating medium pump running until the
temperature drops to a low temperature, e.g. 60oC. (In a real machine, turning off the pump
can be decided in regards to several aspects including use of afterheat in the generator for
extra cooling etc.)
2. Wait for some time to dilute the solution in the generator and shut off the valves between
the generator and absorbers, i.e. V1 and V2 in Figure 6.1. (Shutting off the valves may not be
necessary in a real machine. But it was necessary in the setup because of the overflowing
solution described in section 5.1.4)
3. Turn off the solution pump, chilled water- and cooling water pump. The order of shutting
off these components is not important.
6.5 Conclusions
Since control of the setup had to be done manually, no attempt was made to obtain an
optimized set of control parameters. Therefore although sequences of startup and shutdown
operations were described in the previous section, the information is not enough for an
automatic control. The present phase of the chiller development is too early to provide such
information.
7 Backup systems
Because of the transient and intermittent nature of solar energy, many solar cooling systems,
especially large ones, have backup systems for continuous operation. Most of the backup
systems may be classified into two categories namely, passive- and active backup depending
on its primary source of energy input.
Passive backup systems includes mostly storages, i.e. hot water, cold water and PCM storage
that uses phase changing materials for cooling or heating.
(a) Storage tank (b) gas heater (c) small compression chiller
Figure 7.1 backup systems
Figure 7.1a shows a schematic diagram of a storage tank. Although arrangement of flows,
i.e. design of flow circuits can vary, purpose of the design is to ‘store’ excess energy for later
uses. For higher energy density, present R&D is, for small scale applications, mostly focused
on finding phase-changing materials including ice, salt hydrates, eutectic salt solutions and
etc. It is very encouraging that thermal storage is very active these days from the fact that
such researches would greatly benefit, first of all, solar cooling industry.
On the other hand, active systems include fuel- or electric heaters (e.g. Figure 7.1b), backup
chiller (Figure 7.1c) etc., which uses other source of energy to ensure continuous cooling
supply.
For detailed information on backup systems, please refer to the related literature including
e.g. Henning [2004].
Since a solar system is not designed to operate only in summer, selection of backup systems
should be made also in relation to heating operation. But for only cooling, cold storage is
more advantageous than hot storage because more energy needs be stored in hot storage to
supply the same amount of cooling unless the chiller COP is larger than 1.
For the same reason, use of a backup heater should be avoided for cooling supply with
absorption chillers with COP less than 1. Therefore if an absorption chiller cannot compete
with a compression chiller in terms of primary energy saving, it would be rational to install a
small-capacity compression machine as a backup chiller.
For this reason, if a backup system is required for cold production, a half-effect LiBr
absorption solar cooling system should be equipped with a small compression chiller. And
capacity and control of such a system should be decided for the overall system performance.
Need of such a ‘systematic design’ can be confirmed, for example, by Figure 7.2.
System 1
0.8
System 2
0.6
System 3
I (kW/m2)
Q(kW)
0.2
Radiation
0
0 4 8 12 16 20 24
Time(hr)
Figure 7.2 Cooling load and supply
Figure 7.2 compare an exemplary cooling load profile and cooling supply curves for different
solar cooling systems.
First of all, one should note that peak of solar radiation (solar gain) does not coincide with
that of cooling load. Peak of cooling load is always delayed by a certain time due to the
building’s thermal mass. And it is also assumed that the peak of cooling supply is also behind
the solar radiation in the figure, which suggests that all solar cooling systems in figure 7.2 are
equipped with storages and/or backup chillers.
Regarding the characteristic curves in Figure 7.2, main points in the design of a solar cooling
system are ‘cooling power’ and ‘storage’.
Cooling power will decide the magnitude of the peak of cooling power curve. In Figure 7.2,
the cooling power becomes smaller from System 1 to 3. On the other hand, storage capacity
will decide the capacity and energy consumption of backup systems for cold production,
which is more or less proportional to the area between the cooling load curve and the cooling
demand curve.
Cooling power and storage capacity can be effectively designed to maximize the benefit of
solar cooling. For example, systems 2 and 3 in Figure 7.2 do not produce any excess cooling
during their (solar) operation and they should provide the difference of cooling demand and
supply by backup systems for cold production. On the other hand, System 1 produces excess
cooling during the morning and early afternoon, which is large enough to cover the delayed
cooling demand late in the evening. Therefore if System 1 is equipped with appropriate
storage and control means, maximum benefit would be obtainable.
Although such a design, i.e. over-design and storage, may not be possible for high-cooling
load applications like hotels and multi-story office buildings primarily due to spatial
limitation for solar collectors, it would be a good solution for large houses or small- and
medium offices, which are the main targets of this project.
8 Assessment
Section 8.1 summarizes steady-state simulation models and the results of the proposed
absorption chiller and its components starting from the fundamentals of thermodynamics and
transport phenomena.
Section 8.2 evaluates a solar cooling system mainly from the financial point of view.
This section describes the heat and mass transfer models for all components in the proposed
absorption chiller including modeling strategy, solution methods and component and system
simulation results.
A schematic diagram of the chiller is given in Figure 8.1 with all major state points
numbered for modeling.
34
3
20 23 CON
7 6
33
pl pm
ph 25
16
32 38
MPA
MPE
LPA
36
LPE
GEN
31 37
35
21
1
2 18 20 23 24
8 5 17 14 16
Ref-tank 4
Sol
-tank
Ref-HEX LT-HEX HT-HEX
The system consists of six two-phase heat exchangers, three single-phase heat exchangers
and two mixing tanks connected to each other. Because the system has so many components
coupled to each other providing many interactive parameters to be included in the model, it is
important to develop equations that would facilitate fast convergence of the solution.
Since the governing equations for absorption and desorption processes are highly non-linear,
various iterative methods are used for the solution of simultaneous non-linear equations,
which requires much of extra procedures other than solving the original equations
themselves. For example, the modified Newton-Raphson method by Powell (1970) requires
determination of a Jacobian matrix to correct intermediate solutions at every step until the
solution converges [e.g. Grossman and Michelson (1994) and Zhuo (1995)]. And because
there are always possibilities of physically unrealistic solutions, enough care should be given
to the initial conditions and the intermediate solutions in this approach. In many cases, these
extra measures impose excessive complications and inefficiencies.
Generally, much of the non-linearity in an absorption system originates from the non-linear
nature of the working fluids’ properties. Enthalpy of solution, for example, is often treated as
an implicit function of independent variables because of its high non-linearity. Consequently,
energy equations that inevitably include solution enthalpies are expressed implicitly in terms
of the original independent variables. Then, the corresponding property subroutines should
be called in every time to update the enthalpy terms and therefore slow the convergence.
One way to avoid all these problems regarding the solution of non-linear equations is to
develop linear governing equations in the first place or to transform the non-linear equations
into equivalent equations in more appropriate forms for simple and quick solution. This
approach is adopted here and a great deal of effort is given to the transformation of equations
in the following sections.
Two-dimensional numerical models such as proposed by Yang and Wood (1992) and
Wassenaar (1994) would provide no such problem but it is not a practical approach because
of its excessive requirements in memory and computing time for a system simulation. A
practical alternative may be a one-dimensional differential model based on empirical heat
and mass transfer correlations. Although the number of variables would increase in
proportion to that of differential control volumes and so does the computing time, the heat
and mass transfer models would be as accurate as the empirical heat and mass correlations.
All governing equations will be developed following the directions above. Details of the
modeling are given for each component in the following sections.
8.1.2 Generator
The generator produces refrigerant vapor by heating refrigerant-rich solution with a heating
medium. For a LiBr-water absorption chiller, a falling film type generator is preferred
because the pressure drop in the vapor flow is relatively small. A schematic diagram of a
falling film type generator is given in Figure 8.2.
25
ph T36 T25, m25
16 m16 xb16 xi16
y i=1, z=0
z
mi Tbi xbwi Ti xi mvi
Twi i wi Tvi i
36
Q dmi Tv
Tv h v
dA
Film Vapor
n, z=L
m35 T35 m24 xb24 xi24
24
The generator in Figure 8.2 is a shell & tube type heat exchanger where the heating medium
flows through the shell side in an upward cross flow configuration and refrigerant vapor is
generated from the solution that flows downward inside the vertical tubes.
- The same working conditions are applied to all the tubes regardless of the position
- Heating medium flow is counter-current to the solution flow
- Pressure drop along the vapor flow is negligible
For the ith liquid control volume shown in Figure 8.2, the following governing equations can
be derived.
and the average temperature difference between the vapor and the solution is defined by
1 v
ΔTavg
v
≡
2
( Ti + Ti +v1 − Ti i − Ti +i 1 ) (8.5)
where ΔTlavg denotes the mean temperature difference between the heating medium and the
solution defined by
1 w
ΔTavg
l
≡
2
( Ti + Ti +w1 − Ti b − Ti +b1 ) (8.7)
where K denotes the mass transfer coefficient and Δxavg the mean concentration difference,
which is defined between bulk and interface concentrations by
1 i
Δxavg ≡
2
( xw,i + xwi ,i+1 − xwb ,i − xwb ,i+1 ) (8.9)
For the ith vapor control volume shown in Figure 8.2, the following governing equations can
be derived.
∑ m& =m& v
i − dm& i − m& iv+1 = 0 (8.10)
∑ Q& = ( m& h v v
i i − dm& i h v − m& iv+1hiv+1 ) + hTv dAΔTavg
v
=0 (8.11)
All equations above will be expressed in terms of the primary variables including mass flow
rate m& , and temperature T and concentration x and transformed into appropriate forms for
easy and quick solution.
Note that the arithmetic means are used for the driving potentials in Equation (8.5), (8.7) and
(8.9) instead of logarithmic means. This is valid for infinitesimal differential control volumes
and would provide acceptable accuracy for reasonably sized differential elements. It is
advantageous that they are all expressed linearly in T and x. But it is not certain at this
moment whether Equation (8.8) is linear in independent variables because the interface
concentration xi is not fully described yet.
On the other hand, Equation (8.2) and (8.3) are nonlinear. Especially Equation (8.3) is
complicated because the desorption heat Qi in Equation (8.4) contains the solution enthalpies
that are highly non-linear in T and x. If the enthalpy terms in Equation (8.4) are not to be
expressed explicitly in terms of the primary variables, the desorption heat Qi, being an
implicit function of the variables, presents extra difficulties in the solution process. This is a
significant disadvantage especially for a differential model where many of such equations
should be solved simultaneously.
In the following, these non-linear equations are transformed to appropriate forms for quick
and simple solution using some thermodynamic and heat and mass transfer theories.
Solution of Equation (8.8) requires knowledge of the interface concentrations. The goal of
this section is to express the interface concentration explicitly in terms of the bulk properties
of the falling film, preferably with a linear equation. For this purpose, two famous studies of
Nakoryakov and Grigoreva (1980) and Yüksel and Schlünder (1987) are used as presented in
the following.
Nakoryakov and Grigoreva (1980) analyzed the entrance region of a laminar film flow over a
vertical wall and considered two thermal boundary layers, one of which grows from the wall
and the other from vapor-liquid interface. The two layers eventually meet somewhere down
stream. The idea is illustrated in Figure 8.3.
Tb, xbw
y' y
wall interface
Figure 8.3 Double boundary layers in entrance region
At z=0, a solution of the bulk temperature Tb and concentration xb is supplied. Soon the two
boundary layers begin to grow. The one whose thickness is denoted by δT1 grows from the
wall to the interface and the other δT2 grows in the opposite direction. The two boundary
layers meet downstream at z=zT. The authors assumed a constant velocity u in the layer δT2
and solved the conduction and diffusion equations for temperature and concentration.
∂T k ∂ 2T
u = (8.12)
∂z ρ C p ∂y 2
∂xw ∂2 x
u = D 2w (8.13)
∂z ∂y
T −Ti 1 ρC pu y
= erf (η1 ) η1 = (8.14)
T −T
b i
2 k z
where erf(η) is the error function. And the solution of Equation (8.13) is given similarly by
xw − xwi 1 u y
= erf (η2 ) η2 = (8.15)
xw − xw
b i
2 D z
Since the heat transfer rate at the interface should be equal to the latent heat of vapor
desorbed or absorbed, inserting T and xw in Equation (8.14) and (8.15) into a heat balance
equation at the interface, i.e. k∂T/∂y=-hfgρD∂xw/∂y and rearranging it gives
C p Le
xwb − xwi =
h fg (T i
−Tb ) (8.16)
From Equation (A2.6) in Appendix A2, Ti in Equation (8.16) can be expressed on Celsius
scale as
Inserting Equation (8.17) into Equation (8.16) and rearranging it for xiw gives
−1
⎡ h fg ⎤
where C1 ≡ ⎢ A (T dew + 273.15 ) + ⎥ , C2≡ -C1B and C3≡273.15(C1+C2).
⎢⎣ C p Le0.5 ⎥⎦
Equation (8.16) and thus Equation (8.18) is valid in the entrance region where the two
boundary layers in Figure 8.3 are still developing, i.e. 0<z<zT. Nakoryakov and Grigoreva
(1980) also gave the entrance length zT as
The entrance length zT has been calculated for a falling film flow of 50 % solution at 53 oC.
The result is illustrated against Reynolds number in Figure 8.4.
10
21
16
zT(mm)
zT/δ
4
10
2
5
0
0 400 800 1200
Re(=4Γ /μ )
Figure 8.4 Thermal entrance length in falling film flow (xw=0.5 and T=53 oC)
According to Figure 8.4, zT is only 2 mm for Re=320, which is corresponding to the flow rate
of Γ=0.16 kg/m s. Considering that this Reynolds number is well beyond typical design flow
rates for falling film heat exchangers in absorption machines, Equation (8.18) would be
applicable only to the 1st differential control volume if its size is not selected smaller than 2
mm.
Yüksel and Schlünder (1987) suggested a method to determine interface conditions based on
heat and mass transfer analogy. Average Sherwood numbers determined from the calculated
interface concentrations were in good agreement with those determined from their infrared
measurement technique.
They have derived an interface model for fully developed falling film flows from the heat
and mass transfer analogy [see e.g. Holman(1986)] given by
hT
= ρ sol C psol Le1− n (8.20)
K
where hT and K denote the liquid-side heat transfer coefficient at the interface and the mass
transfer coefficient respectively.
Yüksel and Schlünder (1987) suggested that the exponent ‘n’ should be a value between 0.4
and 0.5 according to the Schmidt number exponents in correlations for isothermal falling film
absorption.
Their study is based on the stagnant film model [see e.g. Bird et al (1965)], which assumes a
thin stagnant film adjacent to the phase-changing surface. The thickness of a stagnant film is
shown as Δ in Figure 8.5.
y
z
δ
Ti ds
v
U∞ x∞w T∞ Tb dz
ΔT u θ
dz
dδ
xbw Δx
dm& = n& ds
Δx xi θ
w dm& y
xiw ΔT Ti y v
u
dm& z
interface n& wall interface
(a) Stagnant film model (b) Falling film flow
Figure 8.5 Stagnant film model and falling film flow
The basic ideas of the stagnant film theory are introduced in the following. Although Yüksel
and Schlünder (1987) described the same ideas on molar basis, mass based expressions are
used in the following to be consistent with the rest of this chapter.
Over the stagnant concentration layer 0≤y≤Δx in Figure 8.5a, mass balance equation is given
for the mass flux n& by
dxw
n& = ρ sol D + (1 − xw ) n& (8.21)
dy
d 2T dT
k 2
− nC
& p , H 2O =0 (8.22)
dy dy
⎛ D ⎞ ⎛ x∞ ⎞
n& = ρ sol ⎜ ⎟ ln ⎜ wi ⎟ (8.23)
⎝ Δ x ⎠ ⎝ xw ⎠
D n&
K≡ = (8.24)
Δ x ρ sol ln ( xw∞ / xwi )
Equation (8.22) and the boundary conditions give the heat flux at y=0 as
& p , H 2O ( T ∞ − T i )
nC
q& y =0 = (8.25)
1 − exp ( nC
& p , H 2O ΔT / k )
Replacing q& y =0 = n& hfg in Equation (8.25) gives the heat transfer coefficient at the interface hT
as
k nC
& p , H 2O
hT ≡ = (8.26)
ΔT ln ⎡1 + C p , H O (T i − T ∞ ) / h fg ⎤
⎣ 2 ⎦
Yüksel and Schlünder (1987) used these stagnant film models for the falling film flows
replacing x∞w and T∞ with xbw and Tb respectively.
Inserting Equation (8.24) and (8.26) into Equation (8.20) and rearranging it finally gives the
equation of Yüksel and Schlünder (1987) as follows.
C pH 2O
⎛x ⎞b C psol Le1−n C pH 2O
⎜ ⎟
w
i
= fg
(T i − T b ) + 1 (8.27)
⎝x ⎠w h
As shown above, Yüksel and Schlünder (1987)’s model is based on the assumption that the
original stagnant film model is valid for the falling film flows. But it is questionable if it is
true.
For this stagnant film model to be applicable to the falling film flows, most of all, the
convective mass flux in the transverse direction in the falling film should be equal to the total
mass flux at the interface. The following proves that it is not true.
Assuming that the total mass of the vapor being absorbed at the interface Figure 8.5b is dm& ,
then the mass flow in the transverse direction dm& y in Figure 8.5b may be expressed in terms
of film thickness δ and the velocity at the interface u as
where, for the film thickness δ, the following general correlation can be used
1/ 3
⎛ν 2 ⎞
δ = a ⎜ ⎟ Reb (8.29)
⎝ g⎠
in which the constant a and the exponent b depend on the pattern of the flow.
dm& y dδ
= 1 − ρu (8.30)
dm& dm&
dm& y u Re dδ
= 1− (8.31)
dm& uavg δ d ( Re )
dm& y ⎛ u ⎞
= 1− b ⎜ ⎟⎟ ≡ Fy (8.32)
dm& ⎜u
⎝ avg ⎠
where the ratio is named Fy for convenience. The ratio Fy is the portion of the transverse
mass flux out of total mass flux at the interface. It is found that Fy =0.5 for laminar flows
from Equation (8.32) with b=1/3 and u/uavg=1.5 from Nusselt (1916). Therefore it is clear
that the original stagnant model substantially overestimates the transverse mass flux.
Considering this inequality of transverse and total mass flux, from Equation (8.32), the mass
transfer coefficient can be redefined as
n&
K≡ (8.33)
ρ sol Fy ln ( xwb / xwi )
C pH 2O Fy
⎛x ⎞ b C psol Le1−n C pH 2O
⎜ ⎟
w
i
= (T i − T b ) + 1 (8.34)
⎝x ⎠ w h fg
In order to make Equation (8.34) more suitable for the present application, expanding the
left-hand side in Taylor series and taking only the first two terms gives
⎛ C psol Le1−n ⎞ i
x −x =⎜
b i
⎟⎟ (T − T ) + Π
b
(8.35)
w ⎜ h fg
w
⎝ ⎠
⎝ h ⎠⎝ Fy ⎠
Note that if the exponent n is 0.5, Equation (8.35) is analogous to Equation (8.16) except for
the last term Π on the right-hand side, which can be treated as a residual in the solution
process because it is relatively small.
In the same way as Equation (8.16) is treated, Equation (8.35) is rearranged for xiw as
−1
⎡ h fg ⎤ h fg
where D1 ≡ ⎢ A (T dew + 273.15 ) + ⎥ , D2 ≡-D1 B and D3 ≡273.15(D1+D2 )- D1 Π.
⎣⎢ C p Le1− n ⎦⎥ C p Le1− n
Using Equation (8.18) and (8.36) for the interface concentrations, Equation (8.8) can now be
expressed as follows.
ρ KdA
m& i − m& i +1 − ⎡⎣ Φ iTi b + Φ i +1Ti +b1 + ( Θi + Θi +1 ) T dew + ( Ω i + Ω i +1 ) ⎤⎦ = 0 (8.37)
2
where Фi=C1, Θi=C2 and Ωi=C3 in Equation (8.18) if the ith node is in the thermal entrance
region or Фi=D1, Θi=D2 and Ωi=D3 in Equation (8.36) if it is not.
Equation (8.3) includes the desorption heat Q& i , which is a highly non-linear function of
several independent variables. Since this non-linearity is the main source of the inefficiency
in the solution process, Equation (8.3) will be transformed to allow a simple and quick
solution.
From the thermodynamic study in Appendix A1, it is shown in Appendix A3 that the
desorption heat can be excellently approximated by Equation (A3.20).
For the ith differential control volume of liquid in Figure 8.2, Equation (A3.20) is rewritten as
where ax=xwi is the gradient ∂(1/Tdew)/∂(1/T) at x=xwi and Tv and Ts denote vapor temperature
and equilibrium temperature at the corresponding bulk concentration and the working
pressure respectively.
Since Equation (A3.20) was derived for saturated solutions, taking account of the non-
equilibrium temperatures of bulk solutions at the inlet and outlet of the differential volume,
Q& i can be written as
Assuming Clp,i+1=Clp,i in Equation (8.39) and replacing dm& i = m& i − m& i +1 from Equation (8.1)
gives
hTv dA v (8.41)
− ( m& 35C p 35 ) (Ti +w1 − Ti w ) − (Ti + Ti+v1 − Ti i − Ti+i 1 ) = 0
2
The last term on the left-hand side of Equation (8.41) is the heat transfer rate from the vapor
flow. The interface temperatures Tii and Tii+1 should be expressed in a similar way to the case
of interface concentrations.
From Equation (8.17) and (8.18), the interface temperature Ti is written for the developing
region as
And similarly for fully developed region, from Equation (8.17) and (8.36), it is
⎛ l hTvdA ⎞ b ⎛ hv dA ⎞
−⎜ mC
& i p,i − Φ′i ⎟Ti + ⎜ m& i+1Clp,i + T Φ′i+1 ⎟Ti+b1 + ⎡⎣ax=xi h fg + Cpv (T v −Ti s ) + Clp,iTi+s1 ⎤⎦ ( m & i+1 )
&i − m
⎝ 2 ⎠ ⎝ 2 ⎠ (8.44)
hvdA
& 35Cp35 ) (Ti+w1 −Ti w ) − T ⎡⎣Ti v + Ti+v1 − ( Θ′i +Θ′i+1 ) T dew − ( Ω′i +Ω′i+1 ) ⎤⎦ = 0
−( m
2
where Ф'i=C4, Θ'i=C5 and Ω'i=C6 in Equation (8.42) if the ith node is in the thermal entrance
region or Ф'i=D4, Θ'i=D5 and Ω'i=D6 in Equation (8.43) if it is not.
∑ Q& = m& ( h v
i i
v
− hiv+1 ) − dm& i ( hv − hiv+1 ) + hTv dAΔTavg
v
hTv dA v (8.45)
= m& iv C pv (Ti v − Ti +v1 ) − dm& i C pv (T v − Ti +v1 ) + ( Ti + Ti +v1 − Ti i − Ti +i 1 ) = 0
2
The temperature of the vapor being generated in the ith control volume can be assumed as
Tv=(Tii+Tii+1)/2. Then Equation (8.45) becomes
+ − − − ⎥ (Ti + Ti +1 ) = 0
i
m
& C
⎜ i p T m
& C
⎟ i ⎜ i +1 p T
⎟ i +1 ⎢ (8.46)
⎝ 2 ⎠ ⎝ 2 ⎠ ⎢⎣ 2 ⎥⎦
The interface temperatures in Equation (8.46) are given in Equation (8.42) and (8.43). Then
Equation (8.46) can be rewritten as
⎛ v v hTv dA ⎞ v ⎛ v v hTv dA ⎞ v
⎜ m& i C p + ⎟ Ti − ⎜ m& i +1C p − ⎟ Ti +1
⎝ 2 ⎠ ⎝ 2 ⎠
(8.47)
⎡ ( m& i − m& i +1 ) C pv + hTv dA ⎤
−⎢ ⎥ ⎡⎣Φ′iTi + Φ′iTi +1 + ( Θ′i + Θ′i +1 ) T + ( Ω′i + Ω′i +1 ) ⎤⎦ = 0
b b dew
⎣⎢ 2 ⎦⎥
where Ф'i=C4, Θ'i=C5 and Ω'i=C6 in Equation (8.42) if the ith node is in the thermal entrance
region or Ф'i=D4, Θ'i=D5 and Ω'i=D6 in Equation (8.43) if it is not.
In the previous sections, five governing equations were derived as summarized in Table 8.1
for each differential element. For a generator with n differential elements, the total number of
equations is thus 5n.
UdA w
Overall heat transfer: w
m& 35C p 35 (Ti+1 − Ti w ) −
2
(Ti + Ti+1w − Tib − Ti+1b ) = 0 Equation
(8.6)
Mass transfer in liquid film:
ρ KdA ρ KdA Equation
m& i − m& i+1 − ⎡⎣Φ iTi b + Φ i +1Ti+1
b
+ ( Θi + Θi +1 ) T dew ⎤⎦ = ( Ωi + Ωi +1 ) (8.37)
2 2
*
Only the bold characters are treated as independent variables in the solution process.
On the other hand, six primary variables, namely Tw, Tb, xb, m& , Tv and Tdew, were identified.
Given the working pressure Tdew, the total number of unknowns is thus 5(n+1) for the
generator with n differential elements. Therefore, the generator can be solved only when the
boundary conditions are given for the rest of the variables. In a system simulation, these
boundary conditions and even the working pressure are determined in the relations with other
components but they are arbitrarily given in this section for illustration of the solution
process. The solution method is briefly introduced in the following.
Given the working pressure, for a generator with n differential elements, 5n equations and 5
boundary conditions form the following matrix equation.
[ A][ B ] = [C ] (8.48)
where [A] is a 5(n+1)×5(n+1) coefficient matrix, [B] is a 5(n+1)×1 column matrix for the
unknowns (solution vector) and [C] a 5(n+1)×1 column matrix for the residuals and
boundary conditions. Although no special matrix solver is required to solve Equation (8.48),
Cholesky’s method has been chosen for its good computational economics (James et al,
1985).
Bk(n+1)+i =
B Twi for k=0 (8.49)
Tbi for k=1
xbi for k=2
m& i for k=3
Tvi for k=4
An augmented coefficient matrix, which is the matrix [A] added with an extra column for the
components ci from [C], is formed using the governing equations and boundary conditions as
follows.
Once matrix A% is complete, a matrix solver subroutine is called and determines the 5n
unknowns.
Iteration of Equation (8.48) is necessary because some of the components in [A] include the
primary variables. Consequently, the matrix [A] should be updated from the previous
intermediate solution and Equation (8.48) should be solved repeatedly until the convergence
criteria are met. A relaxation factor can be used between intermediate solutions to stabilize
the solution process as in
bji=R·bji+(1-R)bj-1i (8.51)
where the ‘bji’ denotes the ith component of the solution vector [B] from the jth iteration step
and ‘R’ is the relaxation factor between 0 to 1.
The number of differential control volumes n should be decided considering accuracy of the
solution and the required memory size and computing time. Near the entrance where
temperature and concentration profiles are developing, finer elements would be necessary to
follow the rapid changes. But such a small element is not necessary downstream, where
profiles are already developed and it would only waste computer memory and would delay
the solution process. Non-uniform grid would be desirable in this situation.
In order to provide a reference, a solution has been attained using 200 evenly sized
differential elements for a 1 m-long generator, i.e. the length of each control volume is 5 mm.
A strongly subcooled 50 wt% solution with 66oC (Ts=74.3oC at 11.5kPa) was supplied at the
top so that rapid changes of temperature and concentration profiles can be clearly seen. The
solution is shown in Figure 8.6.
90 0.56
85
0.54
80
T(oC)
xw
0.52
75
p=11.5 kPa (Tdew=48.6oC)
heating medium 0.5
70 p=11.5 kPa (Tdew=48.6oC)
bulk solution bulk solution
bulk vapor interface
interface
65 0.48
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
z/L z/L
(a) Temperature profiles (b) Concentration profiles
Figure 8.6 Uniform grid solution
Note that the bulk and the interface temperature profiles cross at around z/L=0.03. This
profile crossover means that absorption takes place in the entrance region due to the
subcooling of the supplied solution.
Note also that the profiles change rapidly until z/L=0.2 and become almost linear after that. It
suggests that the fine grids are not necessary downstream. In order to save memory and
computing time, several non-uniform grid systems have been considered. It was found that
an exponential grid could provide economic and accurate solutions. Position of the ith node in
such an exponential grid system is given by
zi exp ⎡⎣ F ( i − 1) ⎤⎦ − 1
= (8.52)
L exp ( Fn ) − 1
where ‘n’ is the total number of control volumes and ‘F’ is an adjustable constant for varying
grid sizes. Figure 8.7 compares an exponential grid solution and the uniform grid solution
from Figure 8.6 on a logarithmic scale.
90 0.56
uniform grid
non-uniform grid
85
0.54
80
T(oC)
xw
0.52
75
0.5
70
uniform grid
non-uniform grid
65 0.48
0.01 0.1 1 0.01 0.1 1
z/L z/L
(a) Temperature profiles (b) Concentration profiles
Figure 8.7 Uniform and non-uniform grid solutions
Grid sizes were chosen for Equation (8.52) with F=0.5 and n=8 for the non-uniform grid
solution in Figure 8.7. The same convergence criterion was applied to both solutions, which
requires that the two consecutive solutions for the bulk concentration at the generator outlet
should be within the absolute difference of 10-6 (i.e. 0.0001 LiBr weight %).
The uniform solution converged after 16 iterations and non-uniform did after 13. The
computing time was 226.31 sec for the uniform solution and 0.11 sec for the non-uniform
solution on an Intel Pentium 4 PC (CPU clock speed 2.4GHz). This particular uniform grid
solution cost 22.3 times more memory and 2057 times more computing time than the non-
uniform grid solution. On the other hand, the differences between two solutions are 3×10-3K
and 0.017 wt% respectively in the bulk temperature and concentration at the generator outlet,
which are negligible in the present application.
23 xb23Tb23 m23
xi23
i=1, z=0
38
mi Tbi xbi T ii
xii i
Twi
dmi Tv hv dA
Qi
Twi+1
Tii+1 i+1
mi+1 Tbi+1 xii+1
xbi+1
21 Film Vapor
37
n, z=L
xi21
38 37 xb21Tb21 m21
Note that in Figure 8.8, the direction of the vapor flow in the ith liquid control volume is
drawn opposite to the actual direction to be consistent with the sign convention used for the
generator.
The MPA is a coiled tube at the top of which a solution distributor is mounted. Solution
flows downward along the outer surface of the coil while cooling water flows spirally
upwards inside the tube coil.
The source of refrigerant vapor, MPE is built around MPA to minimize the distance that the
vapor must travel. Since the bulk vapor flows horizontally into the solution in this
arrangement, the heat transfer between vapor and the solution can be neglected.
From these assumptions, the governing equations of MPA become identical to those of the
generator except for the negligible liquid-vapor heat transfer.
For the ith liquid control volume shown in Figure 8.8, the following governing equations can
be derived.
where ΔTavg denotes the mean temperature difference between the two flows defined by
1 w
ΔTavg ≡
2
(Ti + Ti+w1 − Ti b − Ti+b1 ) (8.58)
where Δxavg denotes the mean concentration difference between bulk and interface defined by
1 i
Δxavg ≡
2
( xw,i + xwi ,i +1 − xwb ,i − xwb ,i +1 ) (8.60)
Note that the governing equations are the same as those for the generator except that the heat
transfer term between liquid and vapor is neglected in Equation (8.55).
Except for Equation (8.55), all the other equations are identical to those of the generator and
therefore the final equations will not be repeated.
Equation (8.55) is transformed in a way similar to the case of the desorption heat in section
8.1.2.
From Equation (A3.21) in Appendix A3, for the ith liquid control volume in Figure 8.8, the
absorption heat Q& i can be written as
where Ts denotes equilibrium temperature at the corresponding bulk concentration and the
working pressure.
Since Equation (8.61) is only valid for saturated solutions, taking account of the non-
equilibrium solution temperatures at the inlet and the outlet and replacing Clp,i+1=Clp,i and
dm& i = m& i − m& i +1 , Q& i becomes
The solution process is the same as for the generator except that there are now 5 primary
variables, namely Tw, Tb, xb, m& and Tdew. The governing equations in their final forms are
summarized in Table 8.2.
UdA w
Overall heat transfer: w
m& 37 C p 37 (Ti+1 − Ti w ) −
2
(Ti + Ti+1w − Tib − Ti+1b ) = 0 Equation
(8.57)
Mass transfer in liquid film:
ρ KdA ρ KdA Equation
m& i − m& i+1 − ⎡⎣Φ iTi b + Φ i +1Ti+1
b
+ ( Θi + Θi +1 ) T dew ⎤⎦ = ( Ωi + Ωi +1 ) (8.37)
2 2
Equation
(8.59)
*
Only the bold characters are treated as independent variables in the solution process.
With a given working pressure Tdew, there are 4(n+1) unknowns and 4n equations for n
control volumes in the MPA. Given 4 boundary conditions, Equation (8.48) is solved with a
4(n+1)×4(n+1) coefficient matrix [A], a 4(n+1)×1 column matrix for residuals and a
4(n+1)×1 column matrix [B] for the variables that is defined for i=1 to n+1 by
Bk(n+1)+i =
B Twi for k=0 (8.64)
Tbi for k=1
xbi for k=2
The MPA has been solved for illustration and the result is given in Figure 8.9.
52 0.56
p=2.08 kPa (Tdew=18.1oC)
cooling medium p=2.08 kPa (Tdew=18.1oC)
bulk solution bulk solution
48 0.54 interface
interface
T(oC)
xw
44 0.52
40 0.5
36 0.48
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
z/L z/L
(a) Temperature profiles (b) Concentration profiles
Figure 8.9 Uniform grid solutions
Like the generator’s case, 200 evenly sized grids were used for this case. A superheated 53.0
wt% solution (Ts=45.3 oC at 2.08 kPa) was assumed at the top of MPA and 37 oC of coolant
inlet temperature was assumed at the bottom.
Supply of the superheated solution resulted in sharp profile changes near the top. As is the
case for the generator, prediction of these rapid changes requires fine grids only in the
entrance region. A non-uniform grid system similar to that of the generator can also provide
practically identical results as shown in Figure 8.10.
52 0.56
uniform grid
non-uniform grid
48 0.54
T(oC)
xw
44 0.52
40 0.5
uniform grid
non-uniform grid
36 0.48
0.01 0.1 1 0.01 0.1 1
z/L z/L
(a) Temperature profiles (b) Concentration profiles
Figure 8.10 Uniform and non-uniform grid solutions
Comparison of the uniform- and non-uniform solution showed results that are comparable to
the case of the generator. The uniform grid solution cost 22.3 times more memory and 1150
times more computing time than the non-uniform grid solution. The differences between two
solutions are 5×10-3K and 1.6×10-4 (0.016 wt%) respectively in the bulk temperature and
concentration at the MPA outlet.
The LPA/MPE unit is a heat exchanger that has an evaporator on one side and an absorber on
the other side. Since the pressure drop must be minimized on both sides, both sides were
designed as falling film type. As shown in Figure 8.11, this component is basically a large
vertical tube with two circular liquid distributors mounted at the top to develop falling film
flows on both sides of the tube wall.
i=1, z=0
Tdewm Tb06 xb20
m06 Tb20
xi20
m20
6 20
i
mwi Twi mi Tbi xbi Ti i
xi i
Tdewm
dmwi Tvm hvm Qi dmi Tvl hvl dA
Tdewm
Ti i+1 i+1
mwi+1 Twi+1 mi+1 Tbi+1 xii+1
xbi+1
pm=p(Tdewm) pl=p(Tdewl)
Figure 8.11 Schematic diagrams of LPA/MPE unit and the control volume
The two vertical falling film flows and the control volumes are shown in Figure 8.11.
Although this component is different from the generator and the MPA by the fact that both
fluids experience phase change, the modeling is essentially the same. The flow in the MPE is
a special case of the falling film flow in the generator and the LPA is the same as the falling
film of the MPA. Modeling of this subject is only as complicated as combining the two
falling film models.
Since the MPA is mounted inside the MPE and the LPE around the LPA, the pressure drops
in the vapor flows and the heat transfer to or from the vapor can be neglected.
For the ith liquid control volumes shown in Figure 8.11, the following governing equations
can be derived.
∑ m& =m& i
w
− dm& iw − m& iw+1 = 0 (8.66)
where Q& i is the heat transfer rate from the LPA to the MPE.
∑ Q& = (m& w w
h + dm& iw hmv − m& iw hiw ) − Q& i = 0
i +1 i +1 (8.69)
where ΔTavg denotes the mean temperature difference between the two flows defined by
1 b
ΔTavg ≡
2
( Ti + Ti +b1 − Ti w − Ti +w1 ) (8.71)
where Δxavg denotes the mean concentration difference between bulk and interface defined by
1 i
Δxavg ≡
2
( xw,i + xwi ,i+1 − xwb ,i − xwb ,i+1 ) (8.73)
∑ Q& = dm& i
w
h fg − hTw dAΔTavg
w
=0 (8.74)
where hwT denotes the liquid-side heat transfer coefficient near the vapor-liquid interface
and ΔTavg
w
denotes the mean temperature difference between bulk and interface defined by
1 w
ΔTavg
w
≡
2
( Ti + Ti +w1 − 2Tmdew ) (8.75)
where, in turn, the Tdewm denotes the interface temperature, i.e. the dew temperature of the
steam at the working pressure of the MPE.
Note that a new heat transfer expression, Equation (8.74) is added to define the relationship
between bulk and interface temperatures in the MPE. This is equivalent to Equation (8.72),
the mass transfer equation of LPA, in respect that it defines the conditions at the vapor-liquid
interface in LPA.
Since all the other equations are analogous to those counterparts in the previous sections or
need no more treatment, the following sections will deal with only the energy balance and
overall heat transfer equation, i.e. Equation (8.68)~(8.70).
UdA b
(m& i +1hi +1 + dm& i hlv − m& i hi ) +
2
(Ti + Ti+b1 − Ti w − Ti+w1 ) = 0 (8.76)
Recalling that the terms in the 1st bracket on the left-hand side of Equation (8.76) is the
absorption heat given by Equation (8.62), Equation (8.76) becomes
2
⎟ Ti +1 − ⎜ m& i C p ,i −
⎠ ⎝
l
2
⎟ Ti −
⎠ 2
( Ti + Ti +w1 )
(8.77)
+ ⎡⎣ ax = xi+1 h + C p ,iTi + C p (Tl v − Ti +s 1 ) ⎤⎦ ( m& i − m& i +1 ) = 0
fg l s v
From Equation (8.65) and (8.69), the evaporation heat Q& i is given by
Note that Equation (8.78) is identical to the expression for the desorption heat Equation
(8.40) with Tb=Tw, ax=xi =1.0, Tv= Tsi and Tsi+1=Tdewm.
⎛ w l UdA ⎞ w ⎛ w l UdA ⎞ w
⎜ m& i +1C p ,i + ⎟ Ti +1 − ⎜ m& i C p ,i − ⎟ Ti
⎝ 2 ⎠ ⎝ 2 ⎠
(8.79)
UdA b
+ ( h fg + C lp ,iTmdew )( m& iw − m& iw+1 ) − ( Ti + Ti +b1 ) = 0
2
Governing equations of the final forms are summarized in Table 8.3 below.
2 ⎠
⎟ Ti+1 − ⎜ m& i C p ,i −
⎝
l
2 ⎠
⎟ Ti −
2
(Ti + Ti+1w ) Equation
(8.77)
+ ⎡⎣ ax = xi+1 h fg + C lp ,iTi s + C pv (Tl v − Ti +s 1 ) ⎤⎦ ( m& i − m& i+1 ) = 0
In Table 8.3, 7 primary variables, namely Tw, Tb, xb, m& , m& iw , Tdewl and Tdewm can be
identified. Excluding the two working pressures, the solution vector [B] in Equation (8.48)
becomes a 5(n+1)×1 column matrix defined for i=1 to n+1 by
Bk(n+1)+i =
B Twi for k=0 (8.80)
Tbi for k=1
xbi for k=2
m& i for k=3
m& iw for k=4
Given 5 boundary conditions and the two working pressures, Equation (8.48) can be solved
with a 5(n+1)×5(n+1) coefficient matrix [A] and a 5(n+1)×1 residual matrix [C].
The LPA/MPE unit has been solved for 200 evenly sized grids for illustration and the result
is given in Figure 8.12.
45 0.58
pLPA=0.84 kPa (Tdew=4.5oC)
pMPE=2.08 kPa, pLPA=0.84 kPa
bulk solution
40 bulk-MPE
bulk-LPA
0.56 interface
interface-LPA
35 interface-MPE
0.54
T(oC)
xw
30
0.52
25
0.5
20
15 0.48
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
z/L z/L
(a) Temperature profiles (b) Concentration profiles in LPA
Figure 8.12 Uniform grid solutions
In Figure 8.12, while a superheated 53 wt% solution with 45 oC (Ts=29.9 oC at 0.84 kPa) is
supplied to the LPA, a subcooled refrigerant with 15 oC is supplied to the MPE at 2.08 kPa
(Tdew=18.1 oC). Consequently, there is a large temperature difference between the two fluids
near the top and the profiles change very rapidly in that region. Profiles crossovers are
observed on both sides near the top, which means that desorption takes place on the LPA side
and condensation takes place on the MPE side.
It was found that the sizes of grids must be chosen with care around the inversion point of the
bulk refrigerant temperature profile (z/L=0.04 in Figure 8.12a). Coarse grids around this
point can yield unrealistic solution or cause the solution to diverge.
A non-uniform grid system that is similar to those used in the previous sections can give
practically identical results as shown in Figure 8.13.
45 0.58
uniform grid uniform grid
non-uniform grid non-uniform grid
40
0.56
35
T(oC)
0.54
xw
30
25
0.52
20
15 0.5
The uniform grid solution cost 22.3 times more memory and 2061 times more computing
time than the non-uniform grid solution. The differences between the two solutions are 3×10-
3
K and 1.0×10-4 (0.01 wt%) respectively in the bulk temperature and concentration at the
LPA outlet.
8.1.5 Condenser
A schematic diagram of condenser is given in Figure 8.14, which shows a single tube coil in
a pressure vessel. Cooling water flows upwards inside the coil and refrigerant vapor
condenses outside the coil to be collected at the bottom.
i=1, z=0
34
34
mi Tbi
Tdew i
Twi
v v
Qi dmi T h dA
Twi+1
Tdew i+1
mi+1 Tbi+1
3
33 Film Vapor
25
ph 33
n, z=L
Tb03 m03
By assuming condensation on a vertical plate, such effects as the mixing and transversal
flows, which are likely to be present in a tube coil condenser, are all lumped into the average
heat transfer coefficient to be used.
For the ith liquid control volume shown in Figure 6.14, the following governing equations can
be derived.
where ΔTavg denotes the mean temperature difference between the two flows defined by
1 w
ΔTavg ≡
2
( Ti + Ti +w1 ) − T dew (8.84)
Note that the temperature difference is defined between the bulk cooling water temperature
and the interface temperature, i.e. Tdew, in Equation (8.84). This is only to follow the
definition of condensation heat transfer coefficient to be used.
Inserting Equation (8.84) into Equation (8.83) and rearranging it gives the overall heat
transfer equation as
⎛ UdA ⎞ w ⎛ UdA ⎞ w
⎜ m& 33C p 33 − ⎟ Ti +1 − ⎜ m& 33C p 33 + ⎟ Ti + UdAT = 0
dew
(8.85)
⎝ 2 ⎠ ⎝ 2 ⎠
Considering the superheat of the vapor and Equation (8.81), Equation (8.82) can be rewritten
as
m& i +1C lpTi +b1 − m& i C lpTi b − m& 33C p 33 (Ti +w1 − Ti w )
(8.86)
+ ( m& i − m& i +1 ) ⎡⎣ h fg + C pv (T v − T dew ) + C lpT dew ⎦⎤ = 0
On the other hand, the bulk condensate temperature Tb can be expressed in terms of the wall
temperature and the saturation temperature as
T b = FT
1
wal
+ (1 − F1 ) T dew (8.87)
where the constant F1=0.68 for laminar vertical film condensation given by Rohsenow (1956)
may be used. The wall temperature, in turn, can be expressed as follows from the definitions
of the heat transfer coefficients.
where the constant F2 is defined as F2≡ (1/hlT+1/hwT +Δt/k)-1(1/hwT +Δt/k) with hlT, hwT and
k/Δt denoting the heat transfer coefficient for condensation, cooling water and the wall
conduction respectively.
T b = ( F1 F2 + 1 − F1 ) T dew + F1 (1 − F2 ) T w (8.89)
As summarized in Table 6.4, three primary variables, i.e. Tw, m& and Tdew, were used in
Equation (8.85) and (8.90) to describe each control volume in the condenser.
⎝ 2 ⎠ ⎝ 2 ⎠ Equation
(8.85)
Energy balance for condensate film:
⎡⎣ m& i +1C lp F1 (1 − F2 ) − m& 33C p 33 ⎤⎦ Ti+1
w
− ⎡⎣ m& i C lp F1 (1 − F2 ) − m& 33C p 33 ⎤⎦ Ti w
Equation
− ( m& i − m& i +1 ) ⎡⎣C pv + C lp ( F1 F2 − F1 ) ⎤⎦ T dew + ⎡⎣ h fg + C pv T v ⎤⎦ ( m& i − m& i+1 ) = 0 (8.90)
*
Only the bold characters are treated as independent variables in the solution process.
Therefore, for n control volumes, there are 2n equations for the total number of unknowns of
2(n+1)+1. Besides the two boundary conditions, i.e. m& 1 = 0. and Twn+1=T33, one more
restriction is required. This extra restriction is given either in terms of the dew temperature
Tdew or the total condensation rate m& n +1 . When Tdew is given, m& n +1 is calculated under the
given Tdew. But in a system simulation, Tdew is calculated to keep the mass balance between
the condenser and the generator, i.e. equality of the total mass condensed and generated
( m& n+1 )CON = ( m& 1 − m& n+1 )GEN . Figure 8.15 shows an example of such a calculation.
50 20
48 16
m(kg/m2hr)
46 12
T(oC)
44 8
p=11.5kPa(Tdew=48.6oC)
42 cooling medium 4
refrigerant interface
refrigerant bulk
40 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
z/L z/L
(a) Temperature profiles (b) Local mass flux
Figure 8.15 Uniform grid solutions (200 even grids)
In Figure 8.15a, the bulk condensate temperature has been calculated from Equation (8.87).
In Figure 8.15b, the mass flux downstream is larger due to the larger temperature difference
between the coolant and the dew temperature because the coolant flows upwards.
Use of non-uniform grid is not particularly advantageous for the condenser because the
temperature profiles are monotonous and gradual in the gradients.
A schematic diagram of the LPE is given in Figure 8.16, which shows a single tube coil with
a refrigerant distributor mounted at the top. Chilled water flows upwards inside the coil and
refrigerant vapor evaporates outside. A tray collects unevaporated refrigerant at the bottom.
Tb07 m07
32
7 i=1, z=0
32
mi Tbi
Tdew i
Twi
dmi Tv hv dA
Qi
w
T i+1
Tdew i+1
mi+1 Tbi+1
Film Vapor
1 31 31
n, z=L
b
T 01 m01
Figure 8.16 Schematic diagram of the LPE and the control volume
For the ith control volume in Figure 8.16, the following governing equations are derived.
where ΔTavg
b
denotes the mean temperature difference between chilled water and refrigerant
defined by
1 w
ΔTavg
b
≡
2
(Ti + Ti+w1 − Ti b − Ti+b1 ) (8.95)
∑ Q& = dm& h i
fg
− hTl dAΔTavg
i
=0 (8.96)
where hlT the liquid-side heat transfer coefficient near the vapor-liquid interface and ΔTavg
i
denotes the mean temperature difference between bulk and interface defined by
1 b
ΔTavg
i
≡
2
( Ti + Ti +b1 − 2T dew ) (8.97)
Since analogous equations have been derived in the previous sections for all the equations
above, only the final equations are given in Table 8.5 of Section 8.6.2.
Three primary variables, i.e. Tw, Tb, m& and Tdew, were used to describe this evaporator
model. For each control volume, Equation (8.92), (8.94) and (8.96) should be solved.
Therefore, for n control volumes, 3n equations are available for the total number of
unknowns of 3(n+1)+1.
*
Only the bold characters are treated as independent variables in the solution process.
Given the working pressure Tdew, the system can be solved with three boundary conditions,
e.g. Twn+1, Tb1, m& 1 . But in a system simulation, Tdew should be determined to keep the mass
balance between the LPE and LPA, i.e. equality of the total mass evaporated and absorbed
( m& 1 − m& n+1 ) LPE = ( m& n+1 − m& 1 ) LPA .
For the purpose of illustration, LPE is solved assuming a superheated refrigerant at 15 oC
supplied at the top under the working pressure of 0.84 kPa (Tdew=4.5 oC). Figure 8.17 shows
the result.
16 30
p=0.84kPa(Tdew=4.5oC)
cold medium
refrigerant interface 25
refrigerant bulk
m(kg/m2hr)
12 20
T(oC)
15
8 10
4 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
z/L z/L
(a) Temperature profiles (b) Local mass flux
Figure 8.17 Uniform grid solutions (200 evenly sized)
Because the supplied refrigerant is significantly superheated, evaporation takes place very
quickly near the top. This is why the mass flux is excessively high in the region in Figure
8.17b.
Very fine grids are required to predict such rapid profile changes as in Figure 8.17. Profiles
downstream are, however, smooth and do not require such fine grids. Non-uniform grids
such as used in the other components would effectively represent this characteristic. Figure
8.18 shows an example of such a non-uniform grid solution.
16 30
p=0.84kPa(Tdew=4.5oC) p=0.84kPa(Tdew=4.5oC)
uniform grid uniform grid
non-uniform grid non-uniform grid
m(kg/m2hr)
12 20
T(oC)
8 10
4
0.01 0.1 1 0.01 0.1 1
z/L z/L
(a) Temperature profiles (b) Local mass flux
This particular non-uniform solution is different from the uniform solution by 0.44 % (e.g.
44W for 10kW cooling) in capacity. But it was absolutely advantageous in the memory and
the computing time. The total number of variables was 603 for the uniform solution and it
was 27 for the non-uniform solution. And while the uniform solution took 49.56 sec to
converge, the non-uniform solution took only 0.0625 sec.
There are three single-phase heat exchangers and two mixing tanks in the system. This
section summarizes the governing equations derived for these components. Figure 8.19
shows a schematic diagram showing the flow network between these components.
17
1 2 21
LT-HEX
5
20
Sol-tank 14
3 18 Ref-tank 8 16
23 HT-HEX
Ref-HEX 4 24
Heat exchangers
The ε-NTU model is used for all single-phase heat exchangers. Figure 8.20 shows a
schematic diagram of a heat exchanger.
Tho m& h
Thi
Tci
m& c
Tco
Figure 8.20 Schematic diagram of heat exchanger
For the case of minimum capacity fluid on hot side, the heat exchanger in Figure 8.20 can be
described by
where ε is given, for the type of heat exchanger, as a function of NTU, which is in turn
function of thermal capacities of the heat exchanging fluids and UA of the heat exchanger
[see e.g. Holman (1986)].
Considering that hot-side fluid is minimum capacity fluid for all three heat exchangers, using
the notations in Figure 8.20, the following equations can be derived.
Mixing tanks
The solution tank in Figure 8.19 receives two solution flows at different temperatures and
concentrations from the absorbers above. For the solution tank, the following equations can
be derived.
In the mixing process, since the concentration difference between the mixing solutions is
rather small, the heat of mixing can be neglected. Then Equation (8.104) can be
approximated by
And for the refrigerant tank in Figure 8.19, the following equations are derived.
Neglecting the flashing of refrigerant entering the refrigerant tank at point 4 in Figure 8.19,
the energy balance equation can be written as
ΣQ& = m& 01T01b + m& 02T02b + m& 04T04 − m& 08T08 = 0 (8.107)
8.1.8 Solution
A series of system simulations has been carried out combining all the equations derived in
the previous sections. Although total number of variables and thus that of equations varies
depending on the grid systems set for the components, solution process is not different from
the solution of a single component.
First of all, in order to evaluate the chiller model developed above, comparison was made
with the experimental data from Ch. 5.
Since the chiller setup showed some unexpected heat and mass transfer characteristics in
several components, from a preliminary comparison, the following decisions were made
regarding the heat transfer correlations used in the simulation.
The falling film heat transfer correlation for horizontal tubes, Equation (C.4) in Appendix C
turned out to give excessively larger values (2-5 times) than measured in MPA. And it
predicted large values also for LPE. For this reason, the measured values were used for MPA
and LPE.
Equation (C.2) predicted reasonably accurate heat transfer coefficient for the condenser and
it was used in the simulation.
The falling film heat transfer correlation for vertical plate, Equation (C.5) in Appendix C
gave reasonably accurate prediction for LPA but not for MPE. So, for LPA, Equation (C.5)
was used but for MPE, the experimental correlation Equation (4.30) was used.
Before describing simulation results, it should be understood that the simulation program, at
the moment of preparing the final report, cannot represent the influences of the problems
reported in Ch.5, i.e. loss of refrigerant and refrigerant-solution mixing. Therefore, accuracy
of the simulation is expected to be poor especially regarding the prediction of COP.
As the first example of the comparison, the influence of solution flow rate control on system
performance was simulated, of which the experimental results were introduced in Section
5.2.8 in Ch. 5.
Figure 8.21 shows two different cases of solution flow rate controls. One is constant flow
rate and the other is constant valve opening, i.e. uncontrolled.
400
360
m(kg/hr)
320
280
Tcool=37oC
Controlled flow
240 Uncontrolled flow
60 70 80 90
T35(oC)
Figure 8.21 Controlled and uncontrolled solution flow rate from the generator
For the two different flow rates in Figure 8.21, simulations were carried out and some of the
results are shown in Figure 8.22.
5 0.4
4
0.3
QLPE(kW)
3
COP
0.2
2
While the experimental and simulated cooling capacities are in reasonable agreement in
Figure 8.22a, the predicted COPs are much higher in Figure 8.22b. This is due to the
refrigerant loss as described in Section 5.1.3.
25
Condensation-exp
Evaporatrion-exp
20 Con.=Eva.-sim
m(kg/hr)
15
10
0
65 70 75 80 85 90
T35( C)
o
Figure 8.23 shows the condensation and evaporation of the refrigerant experimentally
determined from the chiller setup and the refrigerant flow rate from simulation. In the
simulation, the condensed refrigerant flow rate and the total refrigerant evaporation rate were
assumed equal and thus only one curve is shown in the figure.
The simulated curve is in reasonable agreement with the experimental evaporation rate,
which explains the agreement between experimental and simulated cooling capacity curves
in Figure 8.22a. The deviation between the condensation rate and the simulated evaporation
rate explains the big difference in COPs in Figure 8.22b, which again suggests that loss of the
refrigerant as explained in Section 5.1.3 is the main reason for the low experimental COPs in
Figure 8.22b.
Figure 8.24 shows experimental and simulation results for the influence of heating medium
temperature on cooling capacity and COP for different coolant temperatures. The
experimental results shown in the figure are the same as those in Figure 5.21 and 5.22 of
Section 5.2.2.
0.4
4
0.3
QLPE(kW)
3
COP
0.2
2 40oC-exp
40oC-exp
37oC-exp
37oC-exp
33oC-exp
1 0.1 33oC-exp
33oC-sim 33oC-sim
37oC-sim 37oC-sim
40oC-sim 40oC-sim
0 0
60 70 80 90 100 60 70 80 90 100
T35(oC) T35(oC)
(a) Cooling capacity (b) COP
Figure 8.24 Influence of coolant and heating medium temperature
Figure 8.24a shows larger deviations between simulation and experimental results for higher
heating medium temperatures. The main reason is that constant solution flow rates were used
in the simulations while they varied in the actual tests.
The discrepancies between predicted and the measured COPs are comparable to the previous
case and again, the loss of refrigerant should be blamed for the actual low COPs.
Parametric analysis
As already described above and also previously in Ch. 5, the present poor system
performance is firstly, due to the loss of refrigerant and the mixing of refrigerant and
solution. Without those problems, it would have been possible to achieve better performance
close to as predicted in Figure 8.22 and 8.24. Another reason is the components turned out to
be particularly poor in their performance, which includes LPA/MPE and the generator (see
Section 5.4). It would be interesting to see how the improvement on a component would
affect the system performance.
Although there are many other parameters, only UA value of a component was chosen for the
analysis. Figure 8.25 shows the variation of cooling capacity when UA value of a single
component is changed at a time. While UA value of a particular component is varied, the
other components were assumed to have experimentally determined UA values.
5
LPA/MPE
GEN
MPA
4.8 LPE
CON
QLPE(kW)
4.6
4.4
4.2
4
0.8 1.2 1.6 2 2.4
Factor x UA
Figure 8.25 Influence of UA values on cooling capacity
As shown in the figure, the influences of the LPA/MPE and the generator are the largest.
This was already expected from the conclusions in Section 5.4.
Since the results in Figure 8.25 were calculated with only one UA value at a time, it is hard
to see their combined influence. Figure 8.26 shows one example.
For the simulation result in Figure 8.26, 130% of experimental UA values were used only for
the LPA/MPE and the generator.
6
Simulation
Experiment
5
QLPE(kW)
1
65 70 75 80 85 90
T35(oC)
Figure 8.26 Influence of increased UA values for the LPA/MPE and the generator
With 30% increase in UA values of the LPA/MPE and the generator, increase of 15-22% of
cooling capacity has been achieved within the temperature range in the figure. But this does
not necessarily suggest increase in the component size because the experimental heat transfer
coefficients for those components are low enough to expect substantial improvement in heat
transfer.
Finally, the UA values from the original design in Table 5.3 were restored for all components
in the simulation program and the cooling capacity was calculated varying the heating
medium temperature as shown in Figure 8.27.
10
Sim+empirical HTC
Sim + experimental HTC
Experiment
8
QLPE(kW)
0
65 70 75 80 85 90
T35(oC)
Figure 8.27 Potential of increasing the cooling capacity
Two cases of simulation results are shown in Figure 8.27. One was calculated with the
experimental heat transfer coefficients from the chiller test and the other with the empirical
correlations in Appendix C.
Since a solar cooling system using the proposed air-cooled half-effect LiBr absorption chiller
does not exist, it is impossible to evaluate its economic performance with realistic figures.
But it may provide a perspective to compare it with a more popular water-cooled absorption
cooling system with similar design specifications.
Recently in the EU project SACE, Henning and Wiemken (Economic study results report,
2003) developed a simulation tool to assess solar cooling systems. They did an analysis on
solar cooling systems in various locations in Europe. They chose three different types of
cooling load profiles including hotel, office and lecture room and carried out wide-ranging
simulations over different chillers, backup systems in different locations with different design
specifications. Among the system configurations they considered, the simulation results for
two absorption cooling systems have been collected and summarized in the following Table
8.6. For more details, one can refer to the original report.
Table 8.6 Water-cooled solar cooling systems1 for a reference hotel2 in Athens. (Henning and
Wiemken, 2003)
Collector Col. area Specific col. area Heat storage Net col. Backup Primary energy
type (m2) (m2/kWcooling) (hours) efficiency saving
Flat 340 11.6 18 17% elec. chiller 36%
Flat 200 6.8 9.4 22% " 39%
1. Single-effect LiBr/water chiller
2. A free-standing six-storey building having a floor space of 643m2 on each level
Only the hotel simulations in Athens are shown here because the energy consumption for
heating is minimal for these cases so that the influence of cooling can be seen relatively
clearly (see load profiles in the original report).
And also the ‘specific collector area’ is similar to the design specification of this project (see
Table 3.3). Therefore comparison with these two cases should be acceptable.
The reference hotel used in the simulation has roughly 3800m2 of floor area that requires a
large cooling power in summer. Assuming that 1m2 requires 30W of cooling, 114kW
machine is needed to cover 100% of the cooling demand. But for both cases in Table 8.6, a
single-effect 29.4kW (=collector area÷specific collector are) absorption chiller was
commonly assumed. Then arithmetic says that ‘solar cooling’ is designed to provide only
26% of ‘total cooling demand’. It is believed that this low percentage of solar cooling is due
to the limited installation area at the top of the building. In terms of solar cooling percentage,
a solar cooling system which the present project is targeting cannot be directly compared
with the cases in Table 8.6.
It should be noted that both systems have electric compression chillers as a backup. Primary
energy saving is 36 and 39% respectively, which are large savings for a 29.4kW machine
installed at such a large building.
Considering that capacities of the absorption chillers in Table 8.6 were significantly under-
designed for the total cooling demand, it is expected that saving potential of a similar solar
cooling system would be much larger for less demanding applications such as large houses
and small- and medium-size offices this project aims at. For example, if a solar cooling
system is designed to take 50-60% of total cooling demand, it would not be difficult to save
70% of primary energy consumption.
In order to give an idea of how much cost an air-cooled system can save, economic
information on a water-cooled solar absorption cooling system was quoted from Henning
(2004) in Table 8.7.
Table 8.7 Economic comparison of cooling systems for a hotel in Athens (Henning, 2004)
Design specification Reference Water-cooled
system LiBr
1 Collector area (m2) 0 200
2 Heat storage volume (m3) 0 18
3 Backup heater (kW) 18.8 17.9
4 Compression chiller power-nominal (kW) 29.4 -
5 Compression chiller power-nominal (kW) 0 29.4
6 Cooling tower-nominal (kW) 0 74.6
Annual energy balance
7 Total annual electricity consumption(pumps, fans.& etc .kWh) 24,811 10,997
8 Annual electricity consumption -chiller only (kWh) 22,388 -
9 Annual heat for cooling/dehumidification (kWh) 0 103,329
10 Annual heat for heating/humidification (kWh) 14,395 14,395
11 Total annual heat demand (kWh) 14,395 117,724
12 Annual useful heat from solar collector (kWh) 0 103,611
13 Annual net collector efficiency 0 30%
14 Annual chiller COP 3.0 0.65
15 Total annual cold production (kWh) 67,164 67,164
16 Total annual water consumption (m3) 0 368.3
Investment costs
17 Solar collecting system including supporting structure 0 50,000
18 Heat storage unit 0 9,000
19 Additional heat source 7,520 7,160
20 Installation (piping, pump…) 10,000 10,000
21 Compression chiller 11,760 -
22 Thermal chiller 0 11,760
23 Cooling tower 0 2,611
24 Pumps 750 750
25 Planning 3,803 9,928
26 Total investment without funding subsidy 41,833 109,209
27 Funding related to the solar collector 0 25,000
28 Total investment with subsidy 41,833 84,209
Annual cost
29 Capital cost 2,811 4,675
30 Maintenance, inspection 837 1,594
31 Annual electricity cost(consumption) 4,962 2,199
32 Annual electricity cost(peak) 1,000 290
33 Annual heat cost(fossil fuel) 606 565
34 Annual water cost 0 1,105
35 Total annual cost 10,216 10,428
36 Annual extra cost for solar system 0 212
37 Annual operation and maintenance cost 7,405 5,753
Table 8.7 compares a reference system with a water-cooled solar cooling system. The
reference system has a compression chiller with COP 3.0 and the solar cooling system has a
single-effect LiBr absorption chiller with COP 0.65.
The two systems are assumed to yield the same cooling capacity as shown in the row 15 in
the table.
For the water-cooled solar cooling system, €50,000 is assumed for solar collector cost, which
is about 60% of the total investment cost of €84,209. An air-cooled half-effect LiBr
absorption chiller requires the same level of heating medium temperature, it would not be
possible to save any money for this item.
Since an air-cooled system does not need all the costs regarding the cooling tower, the
cooling tower cost of €2,611 in the row 23 can be saved. But this saving will be negated, if
an extra compression chiller is to be used for backup.
It would be reasonable to say that investment for an air-cooled system is similar to that of a
water-cooled system.
But it would be possible to save some from the annual operation and maintenance cost of
€5,753 in the row 37. Maintenance and inspection costs of an air-cooled system should be
comparable to that of the reference system. So €757 can be saved from this item. And the
water cost of €1,105, which is almost 20% of the total operation and maintenance cost
€5,753, can be wholly saved. These two savings sum into 32% of the total operation cost of
the water-cooled solar cooling system. Although this calculation does not include the
difference between the operation costs for different backup systems, it shows the saving
potential of an air-cooled solar cooling system.
9.1 Conclusions
In a generic comparison, it has been concluded that air-cooled half-effect H2O/LiBr chillers
in combination with flat-plate solar collectors are a competitive solution in comparison with
water-cooled single-effect H2O/LiBr chillers. This allows operation of a system under
extreme temperature conditions without a cooling tower that requires large power input.
Experiments with a single stage NH3/H2O sorption system show that at least 100°C heating
medium temperature is required for a rejection temperature of 30°C to obtain an acceptable
performance. This makes this system economically less attractive.
Half-effect cycles can be driven with reasonably low temperatures making them attractive
solutions for solar based applications. Both series and parallel flow cycles show comparable
performance, the main difference being the size of the absorbers. From all the heat
exchangers, the effectiveness of the internal heat exchangers has the largest impact on the
COP.
The experimental half-effect chiller setup shows both a cooling capacity and a COP that
significantly deviate from the values predicted with the developed model and the original
geometrical design. Part of this deviation came from some onsite modifications of the
original design and also partly due to financial restrictions. But some construction problems
most probably have a significant effect on this deviation.
From the experiments, it can be concluded that there is an optimum distribution for both
refrigerant and solution between low and mid pressure levels.
The simulation of the experimental chiller suggests that the main reason for the discrepancies
between predicted and measured COP results from the loss of refrigerant associated with
construction problems.
The simulation results also indicate that the low performance of both low-pressure absorber
(LPA)/mid-pressure evaporator (MPE) heat exchanger and the generator prevent the
experimental setup to show reasonable values of cooling capacity.
9.2 Recommendations
It might be too early to conclude what the position of the half-effect chiller in combination
with flat-plate solar collectors is in comparison with other solar-assisted options. The
experimental results obtained show in some point a lower performance than predicted but a
number of reasons were identified that justifies this discrepancy. The system might be
promising but to take conclusions, more experiments will be necessary with a setup where
the shortcomings of this setup have been solved.
Acknowledgements
The authors would like to first of all thank NOVEM for their support of the here reported
project, especially Gerdi Breembroek that supported the project activities when needed.
Henk de Niet designed and coordinated the implementation of the experimental setup Further
Tjibbe van Dijk, Sam Berkhout, Aad Vincenten, Jasper Ruijgrok and Johan Boender have
assisted us with the construction of the setup and solving problems when they occurred.
Without the support of all these people this work would not be feasible. The authors wish to
acknowledge them all.
Notation
A Area [m2]
D Mass diffusivity [m2/ sec]
Cp Heat capacity [kJ/kg K]
d Diameter [m]
f Fugacity [kPa]
G Molar Gibbs energy [kJ/kmol]
Gi Chemical potential [kJ/kmol species i]
Ga Galilei number [-]
g Gravity constant [m2/s]
H Molar enthalpy [kJ/mol]
h Specific enthalpy [kJ/kg]
hfg Latent heat [kJ/kg]
hT Heat transfer coefficient [kW/m2 K]
I Solar radiation [kW/m2]
k Thermal conductivity [kW/mK]
L,l Length [m]
Le Lewis number [-]
M Mass, molar mass [kg], [kg/kmol]
m Molality [mol solute/kg solvent]
m& Mass flow rate [kg/sec]
Nu Nusselt number [-]
n number of mols [-]
p Pressure [kPa]
Pr Prandtl number [-]
Q,V Volume [m3]
Q& Heat transfer rate [kW]
R Gas constant [kJ/K kmol]
Re Reynolds number [-]
r Diameter [m]
S Molar entropy [kJ/K kmol]
Sh Sherwood number [-]
Sc Schmidt number [-]
T Temperature [oC], [K]
U Overall heat transfer coefficient [kW/m2 K]
W Weight [kg]
x Liquid composition [-]
y Vapor composition [-]
Greek letters
α Circulation ratio [-]
γ Activity coefficient [-]
Γ Mass flow rate per unit perimeter [kg/m sec]
δ Film thickness [m]
ε Effectiveness [-]
η Efficiency [-]
θ Angle [degree]
λ Pitch [m]
μ Viscosity, Chemical potential [Pa sec], [kJ/kmol species i]
ρ Density [m3/kg]
σ Surface tension [N/m]
τ Shear stress [N/m2]
υ Ion dissociation number [-]
φ Osmotic coefficient [-]
Abbreviations
ABS Absorber
AHX Absorber Heat eXchange
CON Condenser
COP Coefficient of Performance
DPHE Double Pump Half-Effect
EVA Evaporator
GAX Generator Absorber heat eXchange
GEN Generator
HE Half-effect
HT-Hex High-temperature heat exchanger
LPA Low-pressure absorber
LPE Low-pressure evaporator
LT-Hex Low-temperature heat exchanger
MPA Mid-pressure absorber
MPE Mid-pressure evaporator
Ref-Hex Refrigerant heat exchanger
Ref-tank Refrigerant tank
SE Single-effect
Sol-tank Solution tank
SPHE Single Pump Half-Effect
References
Behrens, D., Dechema corrosion handbook; corrosive agents and their interaction with
materials, VCH Verlagsgesellschaft, Weinheim (Germany), vol.3, p.110, 1988.
Bird, R.B, Stewart, W.E., Lightfoot, W.E., Transport phenomena, Wiley, New York, 1965
Bourouis, M., Valles, M., Coronas, A., Absorption refrigeration cycle driven by thermal
energy at low temperature, Proc. of Eurotherm Seminar 72, Valencia (Spain), 2002.
Chua HT, Toh HK, Malek A, Ng KC, Srinivasan K. Improved thermodynamic property
fields of LiBr-H2O solution. Int. J. Refrigeration, vol. 23, pp.412-29, 2000
Dao, K., Regenerative absorption cycles with multiple stage absorber, US patent, 5157942,
1992.
Debye P, Hückel E. Zur theorie der elektrlyte. Physikalische Zeitschrift, vol. 24, pp.185-208,
1923
Diehl, J.E., Koppany, C.R., Flooding Velocity Correlation for Gas-Liquid Counterflow in
Vertical Tubes, Chemical Engineering Progress Symposium Series, vol.65, pp.77-83, 1969.
DiGuilio, R.M., Lee, R.J., Jeter, S.M., Teja, A.S., Properties of Lithium Bromide-Water
solutions at High Temperatures and Concentrations – I. Thermal Conductivity, ASHRAE
Transactions, vol.96, pp.702-708, 1990.
Fernández-Seara, J., Vázquez, M., Study and control of the optimal generation temperature in
NH3-H2O absorption refrigeration systems, Applied Thermal Engineering, vol.21, pp.343-
357, 2001.
Feuerecker, G., Scharfe, J., Greiter, I., Frank, C., Alefeld, G., Measurement of
thermophysical properties of LiBr solutions at high temperatures and concentrations, Proc.
Int. absorption Heat Pump Conf. ASME, pp.493-499, 1993.
Fujita, T., Falling liquid films in absorption machines, Int. J. Refrig., vol.16, pp.282-294,
1993.
Grossman, G., Wilke, M., Devault, R.C., Simulation and performance analysis of triple-
effect absorption cycle, ASHRAE Trans., vol.100, pp.452-462, 1994
Hartley, D.E., Murgatroyd, W., Criteria for the break-up of thin film layers flowing
isothermally over solid surfaces, Int. J. Heat Mass Transfer, vol.7, pp.1003-1015, 1964.
Henning H.M. and Wiemken Edo, SACE – Solar Air conditioning in Europe (EU project
NNE5/2001/25), Economic study results report (deliverable no. 4, Work package 3), 2003
(http://www.ocp.tudelft.nl/ev/ res/sace.htm )
Holman, J.P., Heat Transfer, 6th ed., McGraw-Hill Book Company, New York, 1986
Iyoki S, Uemura T. Vapor pressure of water-lithium bromide system and the water-lithium
bromide-zinc bromide-lithium chloride system at high temperatures, Int. J. Refrigeration ,
vol. l12, pp. 278-82,1989
James, M.L., Smith, G.M. and Wolford, J.C., Applied numerical methods for digital
computation, Harper & Row publishers, 1985
Jeter SM, Moran JP, Teja AS. Properties of Lithium Bromide-Water solutions at High
Temperatures and Concentrations – III: Specific Heat. ASHRAE Transactions, vol. 98,
pp.137-49, 1992
Kaita, Y., Simulation results of triple-effect absorption cycles, Int. J. Refrigeration, vol. 25,
pp.999-1007, 2002
Keizer, C., Absorption refrigeration cycles, Ph.D. thesis, Delft University of Technology,
The Netherlands, 1982.
Kim, E., Yoon, J.I., Choi, K.H., Seol, W.S., Heat transfer enhancement with a surfactant on
horizontal bundle tubes of an absorber, Int. J. Heat Mass Transfer, vol.45, pp.735-741, 2002.
Kim, D.S., Machielsen, C.H.M., Evaluation of air-cooled solar absorption cooling systems,
Proc. of ISHPC 2002, Shanghai (China), 2002a.
Kim, D.S., Machielsen, C.H.M., Comparative study on water- and air-cooled solar absorption
chillers, Proc. of EuroSun 2002 ISES Conference, Bologna (Italy), 2002b.
Koehler WJ, Ibele WE, Soltes J, Winter ER. Entropy calculations for lithium bromide
aqueous solutions and approximation equation. ASHRAE Transactions, vol. 93, pp.2379-88,
1987
Kojima, M., Fujita, T., Irie, T., Inoue, N., Development of water-lithium bromide absorption
machine operating below zero degrees, Proc. of 21st IIR Int. Cong. of Ref., Wash. D.C.,
USA, 2003
Lee RJ, DiGuilio RM, Jetter SM, Teja AS. Properties of lithium bromide-water solutions at
high temperatures and concentrations-II: Density and Viscosity. ASHRAE Transactions, vol.
96, pp.709-714, 1990
Lenard JLY, Jeter SM, Teja AS. Properties of Lithium Bromide-Water solutions at High
Temperatures and Concentrations – IV: Vapor Pressure. ASHRAE Transactions, vol. 98,
pp.167-72, 1992
Nakoryakov, V.E., Grigoreva, N.I., Calculation of heat and mass transfer in nonisothermal
absorption on the initial portion of a downflowing film, translated from Teoreticheskie
Osnovy Khimicheskoi Tekhnologii, vol.14, pp.483-488, 1980.
Nusselt, W., Die Oberflachenkondensation des Wasserdampfes, VDI Z., vol. 60, p.541, 1916
Nusselt, W., Der Warmeaustausch zwischen Wand und Wasser im Rohr, forsch. Geb.
Ingenieurwes., vol.2, p.309, 1931
Perry RH, Green DW, Maloney JO. Perry’s Chemical Engineers Handbook; 6th ed., New
York: McGraw-Hill Book Co. 1984.
Powell, M.J.D, A hybrid method for nonlinear equations, Numerical Methods for Nonlinear
Algebraic Equation, Gordon and Breach Science Publisher, London, UK, pp. 87-113, 1970
Robinson RA, McCoach HJ. Osmotic and activity coefficients of lithium bromide and
calcium bromide solutions, J. of the American Chemical Society, vol. 69, p. 2244, 1947
Rogers, J.T., Laminar falling film flow and heat transfer characteristics on horizontal tubes,
Can. J. Chem. Eng., vol.59, pp213-222, 1981
Ruiter JP. Thermodynamic description of mixtures and solutions. Arnhem: Kema Scientific
& Technical Reports 1986;4, ISBN 90-353-0044-0.
SACE – Solar Air conditioning in Europe (EU project NNE5/2001/25), Final report (parts 1
& 2), 2003 (http://www.ocp.tudelft.nl/ev/res/sace.htm)
Schulz, S.C.G, equations of state for the system ammonia-water for use with computers,
Proceedings of 13th International Congress of Refrigeration, pp. 431-436, 1973
Sieder, E.N., Tate, C.E., Heat transfer and pressure drop of liquids in tubes, Ind. Eng. Chem.,
vol.28, p.1429, 1936.
Smith, J.M., Van Ness, H.C., Abbott, M.M., Chemical engineering thermodynamics, 6th
edition, McGraw-Hill, 2001
Wassenaar, R., Simultaneous heat and mass transfer in a horizontal tube absorber, Ph.D.
dissertation, TU Delft, 1994
Wilson, E.E., A basis for rational design of heat transfer apparatus, Journal of Heat Transfer,
vol. 37, pp.47-82, 1915
Yang, R. and Wood, D., A numerical modeling of an absorption process on a liquid falling
film, Solar Energy, vol.48, pp.195-198, 1992
Yao, W., Bjustrom, H., Setterwall, F., Surface tension of lithium bromide solutions with heat
transfer additives, J. Chem. Eng. Data, vol.36, pp.96-98, 1991.
Yüksel, M.L., and Schlünder, E.U., Heat and mass transfer in non-isothermal absorption of
gases in falling liquid films, Part I: Experimental determination of heat and mass transfer
coefficients, Chem. Eng. Process, vol.22, pp.193-202, 1987.
Ziegler B., Trepp, Ch., Equation of state for ammonia-water mixtures, Int. J. Refrigeration,
vol. 7, pp.101-106, 1984
Zhuo, C., Absorption heat transformer with TFE-Pyr as the working pair, Ph.D. dissertation,
TU Delft, 1995
Appendix A
This section presents a summary of thermodynamic theories for electrolytes that provides a
theoretical basis for the further development in Section 4.2. Some of the equations are quoted
directly from literature [Smith et al (2001) and Ruiter (1986)] and the details should be
referred to the references.
Gibbs Energy
Usefulness of the Gibbs energy can be readily realized from the fact that the other
thermodynamic properties can be derived directly from it by
⎛ ∂G ⎞
V =⎜ ⎟ (A1.2)
⎝ ∂p ⎠T , x
⎡ ∂ (G / RT ) ⎤
H = − RT 2 ⎢ (A1.3)
⎣ ∂T ⎦⎥ p , x
⎛ ∂G ⎞
S = −⎜ ⎟ (A1.4)
⎝ ∂T ⎠ p , x
⎛ ∂ 2G ⎞
C p = −T ⎜ 2 ⎟ (A1.5)
⎝ ∂T ⎠ p , x
⎡ ∂ (nG ) ⎤
d (nG ) = (nV )dp − (nS )dT + ∑ ⎢ ⎥ dni (A1.6)
i ⎣ ∂ni ⎦ p ,T , n
j
where the last term is the contribution of composition change whose summation is over all
species present.
A partial molar derivative of a property such as the one inside the parenthesis of the last term
in Equation (A1.6) is called ‘partial molar property’, which is defined by
⎡ ∂ (nM ) ⎤
Mi ≡ ⎢ ⎥ (A1.7)
⎣ ∂ni ⎦ p ,T ,n j
In the special case of Equation (A1.6), the partial molar Gibbs energy in the last term is
called ‘chemical potential’ of a species i, specially denoted as μi which is given by
⎡ ∂ (nG ) ⎤
μi ≡ Gi = ⎢ ⎥ (A1.8)
⎣ ∂ni ⎦ p ,T ,n j
Chemical potential plays a very important role in mixture thermodynamics. Some important
relations regarding chemical potential are given in the following (see Smith et al, 2001 for
details).
It can be derived that, for π phases in a mixture, the chemical potential of a species is equal in
all phases that are in equilibrium, i.e. μ iα = μ iβ = ... = μ iπ .
G = ∑ xi μi (A1.9)
i
The chemical potentials of different species in any single phase of a system satisfy
∑ x dμ
i
i i =0 (A1.10)
Once chemical potentials of all substances are determined, one can readily calculate the
Gibbs energy from Equation (A1.9), from which all the other properties can be derived.
⎡ ∂ ( nVi*ig ) ⎤ ⎡ ∂ ⎛ RT ⎞ ⎤ RT
Vi ig (T , p, yi ) = ⎢ ⎥ =⎢ ⎜n ⎟⎥ = = Vi*ig (T , p ) (A1.11)
∂ni ⎥ ∂n p ⎠ ⎦T , p ,n p
⎣⎢ ⎦T , p ,n j ⎣ i ⎝ j
where the superscript * was used to differentiate properties of species i at its pure state from
those in a mixture.
Each species has its partial pressure in the phase according to Dalton’s law. All the other
partial properties of species i in a mixture at T and p should be based on the same T but on its
partial pressure pi as follows.
⎡ *ig
pi
⎛ ∂Vi *ig ⎞ ⎤
H (T , p, yi ) = H
i
ig *ig
i (T , p ) + ∫ ⎢Vi − T ⎜ ⎟ ⎥ dp = H i (T , p )
*ig
(A1.12)
p ⎢
⎣ ⎝ ∂T ⎠T ⎥⎦
pi
⎛ ∂Vi *ig ⎞
S (T , p, yi ) = S (T , p) − ∫ ⎜
i
ig *ig
i ⎟ dp = Si (T , p) − R ln yi
*ig
(A1.13)
p⎝
∂T ⎠T
Analogous to the definition of Gibbs energy G=H-TS, the partial molar Gibbs energy Gi , i.e.
chemical potential can be written as
μ i ( = Gi ) = H i − TSi (A1.14)
Inserting Equation (A1.12) and (A1.13) into Equation (A1.14) gives the chemical potential of
species i in an ideal gas mixture as
Since, for the same temperature T, Gi*ig (T , p ) can be expressed in reference to a reference
pressure po, a low pressure where ideal gas assumption is valid, as
p
⎛ p⎞
Gi*ig (T , p) = Gi*ig (T , po ) + ∫ Vi *ig dp = Gi*ig (T , po ) + RT ln ⎜ ⎟ (A1.16)
po ⎝ po ⎠
, inserting Equation (A1.16) into Equation (A1.15) gives another expression for μiig as
⎛ yi p ⎞
μiig (T , p, yi ) = Gi*ig (T , po ) + RT ln ⎜ ⎟ (A1.17)
⎝ po ⎠
A gas behaves differently in a real mixture where the ideal gas assumption does not hold.
The chemical potential of a species i in a real gas mixture, μig is defined, similar to Equation
(A1.17), as
⎛ fˆi ⎞
μ ig (T , p, yi ) ≡ Gi* g (T , po ) + RT ln ⎜⎜ ⎟⎟ (A1.18)
⎝ po ⎠
where fˆi is the fugacity of species i in a real gas mixture replacing the partial pressure yip in
Equation (A1.17).
In order to quantify the difference between the properties of a species in ideal- and real gas
mixture, ‘residual property’ is introduced. For example, residual chemical potential μiR is
defined by μiR ≡μig -μiig and noting Gi*ig=Gi*g at a low po, it is given by
⎛ fˆi ⎞
μ iR = μig − μiig = RT ln ⎜⎜ ⎟⎟ (A1.19)
⎝ yi p ⎠
The ions are electrically charged and each behaves as an individual molecule in an electrolyte
solution. Thus, electrolyte solutions require another ideal fluid called ‘ideal dilute solution’.
In the ideal dilute solution, it is assumed that there is no interaction between solute ions.
In the following, electrolyte models are introduced for the case of a binary electrolyte
solution.
First of all, real mol fraction of solute and solvent is different from the stoichiometric mole
fractions because of the ionization of electrolyte. The definition and relationship of mole- and
weight fractions, i.e. x±= real mole fraction, x=stoichiometric mole fraction, xw= weight
fraction and another useful parameter m, the molality, are shown below.
n2 1
x2± = = (A1.20)
υ n1 + n2 1 + mM 2υ
n1 x1 x1w
m= = = (A1.21)
n2 M 2 (1 − x1 ) M 2 (1 − x1w ) M 1
where subscripts 1 and 2 denote solute and solvent respectively, M is molar weight and the
symbol υ in the equations is the number of ions that dissociates from 1 mole of solute.
Solute model
In Equation (A1.22), since the pure solute does not exist as a liquid in standard state, G1*l
cannot be defined. Therefore, instead of using the pure liquid as a basis for the model, the
aforementioned ideal dilute solution is used.
Since Equation (A1.22) also applies to a reference solution at concentration x1±o , the chemical
potential of this reference solution can be expressed in the same way.
Removing G1*l by subtracting Equation (A1.23) from (A1.22) yields a new expression for the
chemical potential as
x1±
μ1id (T , p, x1± ) = μ1id (T , p, x1±o ) + RT ln (A1.24)
x1±o
Ideal solution is approached when a solution is infinitely diluted. Taking lim m→0, lim
mo→0 on Equation (A1.24) and expressing it in molalities yields
⎡ ⎛ x1± ⎞ ⎤
μ1id (T , p, m) = lim ⎢ μ1id (T , p, x1±o ) + RT ln ⎜ ± ⎟⎥
⎝ x1o ⎠ ⎦
±
x →0
1
x1±o → 0 ⎣
⎛m⎞
= lim μ1id (T , p, mo ) + RT ln ⎜ ⎟ (A1.25)
mo → 0
⎝ mo ⎠
⎛m⎞
= G1∞ (T , p) + RT ln ⎜ ⎟
⎝ mo ⎠
where the superscript ‘∞’ denotes a property of infinitely dilute solution. Infinitely dilute
solution is a hypothetical fluid whose properties are defined by the partial molar gradients of
the corresponding solution properties at infinite dilution, i.e. when the solution approaches
pure solvent, as its Gibbs energy is defined by Equation (A1.26). The reference molality mo
is customarily set to 1 mol.
Because Equation (A1.25) is only valid for an ideal solution, for solute in a real solution,
non-ideality is introduced in the form of activity coefficient γ1 as
⎛ γ 1m ⎞
μ1l (T , p, m) = G1∞ (T , p) + RT ln ⎜ ⎟ (A1.27)
⎝ mo ⎠
It should be noted that Equation (A1.25) and (A1.27) are based on the assumption that whole
ions act as a single group regardless of their polarities.
If we consider + and – ions separately, Equation (A1.27) has to apply to each group of the
ions as follows.
⎛ γ 1+ m + ⎞
μ1+ (T , p, m + ) = G1+∞ (T , p, mo+ ) + RT ln ⎜ + ⎟
(A1.28)
⎝ mo ⎠
⎛ γ 1− m − ⎞
μ1− (T , p, m− ) = G1−∞ (T , p, mo− ) + RT ln ⎜ − ⎟
(A1.29)
⎝ mo ⎠
If the solute dissociates at constant temperature and pressure, the Gibbs energy is constant in
the process. Therefore, the sum of Gibbs energies of the dissociated ions should be equal to
that of the solute before dissociation. Then, it can be written as
which also holds for infinitely dilute solution as G1∞ = υ + G1+∞ + υ − G1−∞ .
⎡⎛ γ + m + ⎞υ ⎛ γ − m − ⎞υ ⎤
+ −
μ1 (T , p, m) = G1 (T , p ) + RT ln ⎢⎜ 1 + ⎟ ⎜ 1 − ⎟
l ∞
⎥
⎢⎝ mo ⎠ ⎝ mo ⎠ ⎥
⎣ ⎦ (A1.31)
⎛γ m⎞
±
= G1∞ (T , p) + υ RT ln ⎜ 1
⎟
⎝ mo ⎠
where (γ1±)υ is defined as (γ1±)υ≡ (γ1+)υ+ (γ1-)υ-, which is called ‘mean ionic activity
coefficient’.
Excess chemical potential for solute, the difference between the chemical potentials of solute
in real- and ideal solution, is given by subtracting Equation (A1.25) from (A1.31) as
Solvent model
Then, analogous to the case of liquid mixture model, chemical potential of solvent in a real
solution can be expressed as
where the activity coefficient γ2± is introduced to take in to account of the non-ideal behavior
of solvent.
In order to provide common basis for both solute and solvent, taking lim x2±→1 (lim m→0)
on Equation (A1.34), the chemical potential of solvent becomes
⎡ ⎛ 1 ⎞⎤
= G2*l (T , p) + lim ⎢ RT ln ⎜ ⎟⎥ (A1.35)
m →0
⎣ ⎝ 1 + mM 2υ ⎠ ⎦
= G2*l (T , p) − RTmM 2υ
, where lnγ2± has been separated from the concentration term, making it inefficient for
application. For this reason, a new coefficient is introduced to Equation (A1.35) to represent
non-ideal behavior of solvent as
Then, excess chemical potential of solvent is given by subtracting Equation (A1.35) from
Equation (A1.37) as
In equilibrium, the chemical potential of a species in vapor phase is equal to that of liquid,
i.e. μig= μil. Using the expressions for chemical potentials in the previous sections,
equilibrium conditions are derived for binary electrolyte systems.
Since there is only solvent in vapor phase (yi=1), the fugacity of the species in gas mixture fˆi
in Equation (A1.18) becomes that of pure species f i * , which is defined by
p
⎛ fi* ⎞
RT ln ⎜ ⎟ = ∫ (Vi* g − Vi *ig ) dp (A1.39)
⎝ p⎠ 0
Using the subscript 2 (solvent) for i, Equation (A1.18) and (A1.39) gives
⎛ fˆ2 ⎞ ⎛ p⎞
μ 2g (T , p ) = G2* g (T , po ) + RT ln ⎜⎜ ⎟⎟ + RT ln ⎜ ⎟
⎝ p⎠ ⎝ po ⎠
p p
= G (T , po ) + ∫ (V
*g
2 2
*g
−V 2
*ig
) dp + ∫ V 2
*ig
dp (A1.40)
0 po
p po
= G (T , po ) + ∫ V dp − ∫ V2*ig dp
*g
2 2
*g
0 0
Since the G2*g(T, po) can be expressed in terms of G2*g(T, p*), i.e. the Gibbs energy of the
pure solvent at its saturated pressure, as
po
Since the reference pressure po is assumed low enough for a gas to behave ideally, V2*g is
equal to V2*ig and therefore Equation (A1.42) becomes
Since the Gibbs energy of the pure solvent in liquid phase, i.e. G2*l(T, p) is given by
G (T , p ) = G (T , p ) + ∫ V2*l dp
*l
2
*l
2
*
(A1.44)
p*
μ (T , p, m) = G (T , p ) + ∫ V2*l dp − φ RTmM 2υ
l
2
*l
2
*
(A1.45)
p*
Since μ2g should be equal to μ2l in equilibrium, equating Equation (A1.43) and (A1.45) and
noting that G2*g(T,p*)=G2*l(T,p*) gives the osmotic coefficient φ as
p*
1
φ= ∫ (V − V2*l )dp
*g
(A1.46)
RTmM 2υ
2
p
Equation (A1.46) defines the relationship between the saturated pressure p* of pure solvent at
temperature T and the equilibrium pressure p over the solution at the same temperature in
terms of the osmotic coefficient φ.
From Gibbs-Duhem relation given in Equation (A1.10), it is given below without derivation
that the osmotic coefficient is related to the mean ionic activity coefficient by
d (φ m ) = dm + md ( ln γ 1± ) (A1.47)
Aqueous LiBr solution is a mixture of water and lithium bromide, the brine that has been
most popular in absorption refrigeration industry thanks to its outstanding thermodynamic
characteristics ever since its first introduction to the industry by Carrier in mid 20th century.
Among the two ingredients, water plays two roles. First of all, water is the refrigerant that
releases and absorbs heat at different pressure levels producing cooling effect. Secondly, it is
the solvent. Pure lithium bromide exists as a white crystal with a very strong affinity to
water. The crystal lacks fluidity and has very poor heat and mass transfer characteristics.
Water provides the needed fluidity and thus facilitates transportation of the substance.
LiBr is an extremely hygroscopic white crystalline solid, which is normally synthesized from
the reaction of the hydroxide and hydrobromic acids, a metallic salt with melting point 552
o
C and boiling point 1265 oC. It is highly soluble in water and dissociates into Li+ and Br-
ions when dissolved.
Matching its popularity in the industry, many studies have been carried out on the
thermodynamic properties of the solution and the properties in the working domain of the
conventional machines are well established. But, as more attention is given to the
unconventional cycles for better use of energy [e.g. high-temperature triple effect cycles of
Grossman et al (1994) and Kaita (2002), a refrigeration cycle of Kojima (2002) and low
Among the early thermodynamic studies, the most prominent may be Löwer (1960). It was
the first complete study that presented practically all thermophysical properties of the
solution. Using a Gibbs energy equation, he successfully described the thermodynamic
properties of the solution based on his own experimental data. But the maximum solution
temperature of 130 oC limits the applicable range of this study, which is rather low for the
present standard.
The most famous work is, however, probably McNeely (1979). He developed a Dühring
equation from the extensive collection of equilibrium vapor pressures and calculated the
solution enthalpies for wide ranges of temperature and concentration using Haltenburger
(1939)’s method. It has been very popular in the industry because it is easy to use and also
quite accurate in the working ranges of conventional absorption machines. But in the high
concentration region, his dew temperatures are inconsistent with more recent measurements
and his enthalpy shows a questionable trend in differential heat of dilution as the author
himself mentioned.
In 1987, Herold and Moran (1987) have reproduced McNeely (1979)’s data using a Gibbs
energy equation with a modified Debye-Hückel model (Pitzer, 1973). Based on a statistical
method, they determined coefficients of the Gibbs energy equation using a limited amount of
data known at the time.
In 1993, Feuerecker et al (1994) carried out a study based on their own pressure
measurements for the solutions in the concentration range from 40 to 76 LiBr wt% and
temperatures from 45 to 190 oC. They reported good agreement of equilibrium vapor
pressure with McNeely (1979) for the solutions of concentration below 60 wt.%, but
significant deviations above this concentration. Although this study seems quite reliable in
the high temperature region, it may not be so in the low temperature region where the
differential heat of dilution calculated from their Dühring equation deviates substantially
from the measurement of Lange and Schwartz (1928).
In 2000, Chua et al (2000) developed a set of equations for the solutions in the concentration
range from 0 to 75 wt.% and the temperature range from 0 to 190 oC. For equilibrium
criteria, they collected 11 sets of equilibrium pressure data but finally chose only two data
sets to develop a Dühring equation. They assumed Dühring’s rule is valid in the entire range
and deliberately curve-fitted the Dühring gradients and intercepts of McNeely (1979) and
Feuerecker et al (1994) with high-degree polynomial equations. Although those two studies
may be reliable sources for the working ranges they were chosen for, it is risky to neglect all
the other experimental data measured in the other regions. Besides, the Dühring’s rule may
not be satisfactory in some regions as McNeely (1979) and Haltenburger (1939) mentioned
potential errors due to the constant Dühring gradient in the high concentration region. The
use of the extreme high-degree polynomial functions also raises a question on their choice of
fitting parameter.
One year later in 2001, Kaita (2001) suggested a new set of equations for the high
temperature and pressure ranges of triple-effect machines by supplementing the vapor
pressure data of Feuerecker et al (1994) with that of Lenard et al (1992) for the high
temperature range and McNeely (1979) for the low temperature range. His results are valid in
the concentration range from 40 to 65 wt% and the temperature range from 20 oC to 210 oC.
He developed a 2nd-degree dew temperature equation to cover the wide pressure range but
did not use it in his enthalpy calculation. Consequently, his dew temperature equation is
inconsistent with his enthalpy.
As described above, all the preceding studies were either limited to narrow working ranges or
failed to provide a simple and accurate description for the solutions in wide working ranges.
The study presented in this chapter is intended to develop a Gibbs energy description for the
solutions in wide ranges of temperature and concentration because, until the present moment,
there is no study that accurately and consistently describes the thermodynamic properties of
the solutions in wide working ranges on a sound thermodynamic basis.
The Gibbs energy concept is a very useful tool in describing thermodynamic characteristics
of mixtures and solutions. The concept has been successfully applied to some working pairs
in the absorption field. It provides a thermodynamically sound basis for the design and
analysis of absorption systems in simple but very effective ways.
Besides Löwer (1960) and Herold and Moran (1987) on LiBr-H2O, Schulz (1973) and
Ziegler and Trepp (1984) also reported successful applications of the concept on NH3-H2O
pair by. Schulz (1973) applied the concept to ammonia-water mixture and provided
correlations valid up to 25 bar. Ziegler and Trepp (1984) modified it and extended the range
of the equations to 500K and 50bar for heat pump applications.
The purpose of the study reported in this section is to provide a Gibbs energy expression
from which all the secondary properties can be derived. In the following, derivation of a
Gibbs energy equation is presented step by step based on the Gibbs energy theory for
electrolytes summarized above.
LiBr aqueous solution is an electrolyte and there are only two species, i.e. LiBr and water in
the solution. In the following, LiBr is denoted by subscript 1 (=solute) and water by 2
(=solvent).
G l = x1μ1l + (1 − x1 ) μ 2l (A1.49)
⎡ ⎛m⎞ ⎤
G l = x1G1∞ + (1 − x1 ) G2*l + x1υ RT ⎢ ln ⎜ ⎟ − 1⎥ + G E (A1.50)
⎣ ⎝ mo ⎠ ⎦
The excess Gibbs energy can also be expressed in terms of osmotic coefficient only by
inserting Equation (A1.48) into Equation (A1.51) as follows.
G = x1υ RT ∫
E
m
(φ − 1)dm (A1.52)
0
m
In the following, based on the electrolyte models presented above, a Gibbs energy equation is
developed for the solution using the available experimental data in the literature.
Below is shown how φ″ and φ′ can be determined from the relevant experimental data.
Solution density
Molar volume of the solution is derived from the Gibbs energy by differentiating Equation
(A1.50) according to Equation (A1.2) as follows.
V1∞ and V2*l are the molar volume of the infinitely dilute solution and that of pure water
respectively. The last term VE is the excess volume, which is defined by
⎛ ∂G E ⎞
VE =⎜ ⎟ (A1.55)
⎝ ∂p ⎠T , xi
m 1 ⎛ ∂φ ⎞
V E = x1υ RT ∫ ⎜ ⎟ dm (A1.56)
0 m ⎝ ∂p ⎠T , x
which makes the relationship between the excess volume and the pressure gradient of
osmotic coefficient clear. So, once VE is given as a function, the function for ∂φ/∂p can be
derived from Equation (A1.56).
Using the solution density data from International Critical Table (1928, hereafter ICT),
Löwer (1960) and Lee et al (1990) for Vl and the pure water data from Schmidt (1979) for
V2*l in Equation (A1.57), V1∞ and VE have been correlated by
2 2
x1V1∞ + V E = x1 RT ∑ bi mi / 2 where bi = ∑ bijT − j (A1.57)
i =0 j =0
The coefficients bij are given in Table 3.1 of Ch.3. Among the several fitting parameters
considered, m1/2 turned out to be the best, which is also suggested by the theory of Debye-
Hückel (1923).
Since V1∞ is not a function of concentration, it is clear that V1∞ =RTb0 from Equation (A1.57).
And then VE is given by
2
V E = x1 RT ∑ bi mi / 2 (A1.58)
i =1
p 2
φ ′′ = ∑
2υ i =1
i ⋅ bi mi / 2 (A1.59)
Using a polynomial function developed for the volume of saturated water V2*l from Schmidt
(1979), Equation (A1.54) can be rewritten as
2 2
V l = x1 RT ∑ bi mi / 2 + (1 − x1 ) R ∑ e jT j (A1.60)
i =1 j =1
Equation (A1.60) reproduces the solution density of ICT (1928), Löwer (1960) and Lee et al
(1990) within the standard deviation of 0.13 %, 0.34 % and 0.38 % respectively (see Figure
A1.1 and A1.2). The overall standard deviation is 0.29%.
2 Lower (1960)
ICT (1928)
0
Lee et al (1990)
150o C
-2
(ρ -ρ cal)/ρ (%)
0
100oC
-2
0
50oC
-2
0 0.2 0.4 0.6 0.8
LiBr weight fraction
Figure A1.1 Illustration of errors in calculated solution density
2000
x1w=70%
1800 65
Density(kg/m3)
1600
1400
40
35
1200 30
20 Lower (1960)
ICT (1928)
10
1000 0
Lee et al (1990)
claculated
p* s
1 p 2
φ′ =
RTυ mM 2 ∫ (V2* g − V2*l )dp − ∑
2υ i =1
i ⋅ bi mi / 2 (A1.61)
p
For ease of calculation, the following equation has been developed using the steam data from
Schmidt (1979).
2 2
RT
V2* g -V2*l = tanh [α - β ln( p ) ] where α = ∑ α jT − j , β = ∑ β jT − j (A1.62)
P j =0 j =0
Since V2*l is negligibly small, the term ‘tanh[α-βln(p)]’ in Equation (A1.62) can be
considered as the compressibility factor with the maximum error of 0.11% in steam volume
up to 1,200 kPa and 270 oC.
1 ⎡ cosh (α − β ln p ) ⎤ p 2
φ′ = ln ⎢ ⎥ − ∑ i ⋅ bi mi / 2 (A1.63)
υ mM 2 B ⎢ cosh (α − β ln p ) ⎥ 2υ i =1
*
⎣ ⎦
In order to calculate φ′ from Equation (A1.63), the saturated steam pressure from Schmidt
(1979) and Perry et al (1984) were used for p* and 6 sets of equilibrium vapor pressures were
collected from the literature for p. φ′ calculated from Equation (A1.63) for the vapor
pressures were fitted by
6 2
φ ′ = 1 + ∑ ai mi / 2 where ai = ∑ aijT − j (A1.64)
i =1 j =0
The coefficients aij are given in Table 3.1 of Ch.3 and some fitting curves are shown in
Figure A1.3.
20oC
4
100 oC
3
φ'
2 180 oC
The fitting results are summarized for each of the data sources in Table A1.1.
It should be noted that some of the data from the data sources were not used. The vapor
pressures from ICT (1928) were found exceptionally higher than the others for the solutions
above 45 wt%. They were disregarded because they are probably in error as McNeely (1979)
and Koehler et al (1987) have suggested. McNeely (1979) reported that his vapor pressures
might be in error above 64 wt% referring to the abnormal trends in differential heat of
dilution in this region and therefore they were disregarded. Also his data below 45 wt% were
also disregarded because it turned out that they are different from ICT (1928) only by the
fitting errors McNeely (1979) introduced in fixing his dew temperature lines in this region.
Inserting Equation (A1.64) and (A1.59) into Equation (A1.53) completes the expression for φ
as
6
ibi
φ = 1 + ∑ (ai + p )mi / 2 (A1.65)
i =1 2υ
Inserting Equation (A1.65) into Equation (A1.48) and (A1.52) gives the expressions for lnγ±
and GE respectively as
6
2 ib
ln γ ± = ∑ (1 + )(ai + i p )mi / 2 (A1.66)
i =1 i 2υ
6
⎡2 ib p ⎤
G E = x1υ RT ∑ ⎢ ( ai + i ) ⎥mi / 2 (A1.67)
i =1 ⎣ i 2υ ⎦
Experimental values of φ and lnγ± for the solutions at 25 oC from Hamer and Wu (1972) and
Robinson and McCoach (1947) are shown with the calculated values from Equation (A1.65)
and (A1.66) in Figure A1.4.
7 T=25oC
Hamer & Wu (1972)
6 Robinson & McCoach (1947)
this study
5
4
φ and lnγ 6
3
φ
2
1
ln γ 6
0
-1
0 0.04 0.08 0.12
m1/2
⎧1 ⎫
p =exp ⎨ ⎡α - ln (θ + θ 2 - 1) ⎤ ⎬ (A1.68)
⎩β ⎣ ⎦⎭
where θ=cosh[α-βln(p*)]exp(φυmM2β) and the coefficients α and β are the same as given in
Equation (A1.62).
Strictly speaking, calculation of p from Equation (A1.68) needs iteration because φ includes
p as in Equation (A1.65). But, since the pressure dependence of φ is negligibly small, the
pressure term in Equation (A1.65) can be safely neglected so that the iteration is not
necessary for calculation of p from Equation (A1.68). The equilibrium pressures calculated in
this way have been compared with the original vapor pressures. The results are illustrated in
Figure A1.5 and the deviations are summarized in Table A1.1.
180oC
10
0
-10
10 130oC
0
-10
(p-pcal)/p(%)
10 100oC
0
-10
10
60oC
0
-10
10 20oC
0
-10
For the concentrations below 40 wt.%, the deviations from Löwer (1960) and ICT (1928) are
negligibly small as can be seen in Figure A1.5. However, above this concentration, Löwer
(1960)’s pressures are substantially higher than those obtained from Equation (A1.68). The
standard deviation of 3.3 % is mainly attributed to the discrepancy in this region.
Agreement with McNeely (1979) is good for the concentrations between 45 and 60 wt.%.
Above 60 wt.%, the deviation becomes larger especially near crystallization limits. Standard
deviation is 1.8% in pressure and 0.29 K in boiling temperature.
Standard deviation from Feuerecker et al (1994) is 1.4 % in pressure and 0.31 K in boiling
temperature, the best consistency among the original data sets used in this study.
The high-pressure data of Lenard et al (1992) agree well with the calculated values from
Equation (A1.68) and also with Feuerecker et al (1994)’s data up to 55 wt.%. The deviation
in this region is 1.8 %. But the difference becomes large above 60 wt.%, to which the overall
standard deviation of 3.5% is mainly attributed.
Consistency with Iyoki and Uemura (1989) is the worst, resulting in the standard deviation of
4.9 %. This large deviation is partly due to the fact that the consistency within the data set is
not much better. They reported a mean deviation of 2.33 % (3.3 % standard deviation) from
the equation of Uemura and Hasaba (1964).
Enthalpy of solution
From Equation (A1.3) and (A1.50), the molar enthalpy of solution is given by
Using the data for pure water from Schmidt (1979), molar enthalpy of pure water H2*l can be
calculated by
p ⎡
T
⎛ ∂V *l ⎞ ⎤
H 2*l = H 2*ol + ∫ C *pl2 dT + ∫ ⎢V2*l − T ⎜ 2 ⎟ ⎥dp (A1.70)
To po ⎢
⎣ ⎝ ∂T ⎠ p ⎥⎦
2 2
, where C lp 2 = R∑ d jT j and V2l = R ∑ e jT j . The coefficients dj and ej are listed in Table 3.1
j =0 j =0
of Ch.3.
∂ ⎛ GE ⎞ 6
2 ∂ai i ∂bi
H = − RT ⎟ = − x1υ RT ∑ ( +
E 2 2
⎜ p )mi / 2 (A1.71)
∂T ⎝ RT ⎠ p , x i =1 i ∂T 2υ ∂T
The enthalpy of infinitely dilute solution H1∞ can be determined using either experimental
heat of solution or heat capacity of the solution. Since experimental heat of solution is only
available for a few temperatures, the heat capacity data of a reference solution were used to
determine H1∞.
For the reference solution of concentration x1o, Equation (A1.69) can be rearranged for H1∞
as
H ol − (1 − x1o ) H 2*l − H oE
H1∞ = (A1.72)
x1o
Since H2*l and HE are given in Equation (A1.70) and (A1.71), H1∞ can be determined from
Equation (A1.72) if the enthalpy of the reference solution Hol is given. For this purpose, as
recommended by Jeter et al (1992), 60 wt% solution was chosen as the reference solution.
Most experimental heat capacity sources agree with each other at this concentration. The
solution heat capacities from Löwer (1960), Feuerecker et al (1994), Jeter et al (1992) and
Rockenfeller (1987) were fitted with a polynomial equation as illustrated in Figure A1.6 to
allow for the calculation of Hol.
T ⎡ ⎛ ∂V ∞ ⎞ ⎤
H1∞ = H1∞o + ∫ Cp∞1dT + ⎢V1∞ − T ⎜ 1 ⎟ ⎥ ( p − po* ) (A1.73)
To ⎢⎣ ⎝ ∂T ⎠ p ⎥⎦
2
C p∞1 = RT −2 ∑ c jT − j (A1.74)
j =0
2.1
2.0
C p(kJ/kg)
1.9
60 wt.%
Lower (1960)
1.8 Feuerecker et al (1994)
Rockenfeller (1987)
Jeter et al (1992)
this study
1.7
0 40 80 120 160 200
T( C)
o
Using Equation (A1.70), (A1.71) and (A1.73), the enthalpy has been calculated from
Equation (A1.69). The zero enthalpies were chosen for pure water and 50 wt% solution at 0
o
C. Figure A1.7 shows some of the results in comparison with literature values. The enthalpy
of this study is relatively in good agreement with Feuerecker et al (1994) and Kaita (2001).
The discrepancy between this study and Feuerecker et al (1994) is believed to have come
from the disagreement in equilibrium vapor pressure that also caused the disagreement in the
differential heat of dilution in Figure A1.8. Seeing that the differential heat of dilution
calculated from their Dühring equation deviates substantially from the experimental data of
Lange and Schwartz (1928), it is questionable whether their Dühring equation can be safely
extrapolated to the low temperature region.
20
180oC
0
-20
20
130oC
0
hlit-hcal(kJ/kg) -20
20
100oC
0
-20
20
60oC
0
-20
20
20oC
0
-20
0 0.2 0.4 0.6 0.8
LiBr weight fraction
Figure A1.7 Comparison of calculated enthalpy with literature values [○: Löwer (1960), △: Chua
et al (2000), □: McNeely (1979), ●: Feuerecker et al (1994), ■: Herold and Moran (1987), ▲: Kaita (2001)]
On the other hand, the deviation from Kaita (2001) has nothing to do with equilibrium vapor
pressure because he did not use his equilibrium equation in enthalpy calculation. Instead, he
used differential heat of dilution data from several sources and the heat capacity data from
Rockenfeller (1987). The difference between the heat capacity of Rockenfeller (1987) and
that of this study can be seen in Figure A1.9. Compared to the heat capacity of this study, the
average heat capacity of Rockenfeller (1987) is smaller below 60 wt% but slightly larger
above this concentration, which probably caused a similar trend in enthalpy difference
between this study and Kaita (2001).
T=25 oC
Lange and Schwartz (1928)
400 McNeely (1979)
Feuerecker et al (1994)
Lower (1960)
hd(kJ/kg water)
Kaita (2001)
300 this study
200
100
The enthalpy of McNeely (1979), Herold and Moran (1987) and Chua et al (2000) show
similar trends against this study for the solutions below 60 wt% as shown in Figure A1.7.
This similarity originates from that fact that all these studies were based on the dew
temperatures of McNeely (1979). Chua et al (2000) deviates from the rest of the group above
60 wt% because they used the vapor pressure of Feuerecker et al (1994) in that region. In the
high concentration region, McNeely (1979)’s enthalpy is smaller than the others because his
dew temperature gradient is much smaller in this region. The second reason is that McNeely
(1979) calculated his reference solution enthalpy using Löwer (1960)’s heat capacity for 50
wt% solution, which is constant in the high temperature region. This is contradictory to the
measurements of Jeter et al (1992) and Rockenfeller (1987), which show appreciable
temperature dependences throughout the whole temperature range. Consequently, the
temperature gradient of McNeely (1979)’s heat capacity curve in the high temperature region
is zero at around 50 wt% and becomes even negative as the concentration increases as shown
in Figure A1.9.
4.4
0%
4 10%
3.6
3.5
20%
3.3
3.1
3.1
30%
2.9
2.7
2.5
Cp(kJ/kgK)
2.7 40%
2.5
2.3
2.1
2.4 50%
2.2
2
1.8
2.2
60%
2
1.8
1.6
2
1.8
70%
1.6
1.4
Figure A1.9. Comparison of calculated heat capacity with literature values [○: Löwer (1960), △:
Jeter et al (1992), □: Rockenfeller (1987), ----: McNeely (1979), __ _ __: Feuerecker et al (1994), ____: this study:]
Speaking of the enthalpy deviations quantitatively, this study is different from Feuerecker et
al (1994) and Kaita (2001) by maximum 6 and 10 kJ/kg respectively. The maximum
differences from McNeely (1979) is as large as 15 kJ/kg, which is roughly equivalent to the
heat transfer rate of a single-phase heat exchanger where 1 kg of 55 wt% solution undergoes
7 K temperature difference.
Entropy of solution
From Equation (A1.4) and (A1.50), the molar entropy of the solution is given by
⎡ ⎛m⎞ ⎤
S l = x1S1(∞T , p ) + (1 − x1 ) S 2(*l T , p ) − x1υ R ⎢ln ⎜ ⎟ − 1⎥ + S(ET , p , m ) (A1.75)
⎣ ⎝ mo ⎠ ⎦
where S∞1 and S*l2 are the molar entropy of the infinitely dilute solution and that of pure
water respectively. The third term is the entropy generation in an ideal mixing and the last
term SE is the extra entropy generation in a real mixing process.
T
C *pl2 p
⎛ ∂V *l ⎞
S 2*l = S 2*ol + ∫ dT − ∫ ⎜ 2 ⎟ dp (A1.76)
To
T po ⎝
∂T ⎠ p
H E − GE 6
2⎡ ib ∂a i ∂bi ⎤ i / 2
SE = = − x1υ R∑ ⎢ ai + i p + T ( i + p) m (A1.77)
T i =1 i ⎣ 2υ ∂T 2υ ∂T ⎦⎥
S1∞ can be calculated by the following equation using Cp1∞ from Equation (A1.74) and V1∞
from Equation (A1.57).
T
C p∞1 p
⎛ ∂V ∞ ⎞
S1∞ = S1∞o + ∫ dT − ∫ ⎜ 1 ⎟ dp (A1.78)
To
T po ⎝
∂T ⎠ p
The zero entropy states have been chosen for pure water and 50 wt% solution at 0 oC. The
results are illustrated in Figure A1.10 with those of Löwer (1960), Feuerecker et al (1994),
Chua et al (2000), Kaita (2001) and Koehler et al (1987).
Except for Koehler et al (1987) that did not present enthalpy data, the other sources show
similar trends in comparison with this study as for the enthalpy.
25 180 C
o
0
-25
130oC
25
0
-25
slit-scal(J/kgK)
100oC
25
0
-25
60oC
25
0
-25
20oC
25
0
-25
Conclusions
Thermodynamic properties of LiBr aqueous solution have been successfully described with a
Gibbs energy equation for the solutions in the concentration range from 0 to 70 wt% and the
temperature range from 0 to 210 oC.
The equation for osmotic coefficient developed in this study is able to reproduce the original
solution density and the equilibrium vapor pressure from the literature within a standard
deviation of 0.29 % and 2.9 % respectively. Solution enthalpy has been calculated using
experimental heat capacity and differential heat of dilution based on the osmotic coefficient
equation. The calculated enthalpy is consistent with the experimental differential heat of
dilution and heat capacity data taken from the literature. This study agree relatively well with
Feuerecker et al [9] and Kaita [12] but deviates substantially from those based on the
equilibrium vapor pressures of McNeely [6] especially in the high concentration and
temperature regions.
The approach adopted in this study has proved its high flexibility and accuracy for describing
electrolyte solutions over wide working ranges by allowing all relevant parameters to be
systematically expressed in a single Gibbs energy equation that can be readily expandable to
other thermodynamic property equations.
Being able to describe the properties of solutions from pure solvent to the highly
concentrated solutions near crystallization limits at temperatures from freezing points to 210
o
C, this study provides a consistent and reliable basis for simulation and analysis of LiBr
absorption systems.
Although the equilibrium conditions for LiBr-Water system has been completely described
by osmotic coefficient φ in Appendix A1, the equation is not easy to handle for modeling
transfer processes. For this reason, an approximate equilibrium equation has been derived
from the original osmotic coefficient equation, which is presented in the following.
Assuming that steam is ideal gas for low pressures, from Equation (A1.46), the osmotic
coefficient φ can be written as
1 ⎛ p* ⎞
φ= ln ⎜ ⎟ (A2.1)
υ mM H 2O ⎝ p ⎠
Since the saturated steam pressures, p* and p can be approximated by lnp*= c1+c2/T and lnp=
c1+c2/T dew within a moderate pressure range, Equation (A2.1) can be rewritten as
1 1
dew
= − φ c2−1υ mM H 2O (A2.2)
T T
Since φ can be approximated by the following linear equation of 1/T within a moderate
temperature range
⎛ ∂φ ⎞ ⎛ 1 1 ⎞
φ =⎜ ⎟ ⎜ − ⎟ + φo (A2.3)
⎝ ∂1/ T ⎠To ⎝ T To ⎠
1 a
dew
= +b (A2.4)
T T
where ‘a’ and ‘b’ are functions of concentration only. For the purpose of illustration, their
average values over the temperature range of 0 to 210 oC are shown in Figure A2.1.
x10 -5
1.25 10
1.2
8
1.15
6
a 1.1 b
4
1.05
2
1
0.95 0
0 0.2 0.4 0.6 0 0.2 0.4 0.6
xw xw
As it is clear from Figure A2.1, both a and b are non-linear functions of concentration. These
functions can also be approximated by a linear equation for a small concentration range.
Since the absolute value of b is very small, it can be set to a certain constant value, e.g.
b=b(xwo), within the small concentration range. Approximating the coefficient a with a linear
equation of the concentration xw, Equation (A2.4) can be rewritten as
1 ⎡⎛ da ⎞ ⎤1
dew
= ⎢⎜ ⎟ ( xw − xwo ) + axo ⎥ + bxo (A2.5)
T ⎣⎢⎝ dx ⎠ xo ⎦⎥ T
( da / dx ) x ⎛ axo ⎞
where A ≡ , B≡⎜
⎜ 1 − bx T ⎟⎟
o
.
1 − bxo T dew
⎝ o
dew
⎠
Equation (A2.6) is useful for modeling desorption and absorption processes because all the
variables involved in the phase equilibrium are explicitly expressed in it. Applications of the
equation can be found in Ch. 5.
As shown in Appendix A1, the enthalpy of a LiBr solution is a complicated function of the
temperature, concentration and pressure. Because the heat involved in absorption or
desorption process is expressed in terms of the solution enthalpies, it would be convenient if
the process heat could be represented by a simple explicit equation directly in terms of the
independent variable.
Desorption is a general term for the regeneration of gaseous species out of a liquid or solid
mixture regardless of how it is done. But in the field of absorption cooling, it means
generally boiling of a solution. Boiling of a solution is different from that of a pure substance
by the fact that the compositions of species vary during the process. Figure A3.1 shows the
desorption process in the p-T-x domain. Several state points are numbered for reference.
Tdew x=0
3'
T2
M1 T1 x1
x1
1 H1
1' x2
Mv Hv4
Q
4
3 1 4 2 2 H2
T3
M2 T2 x2
T3 T1 T2 Tsol
Figure A3.1. Desorption process in p-T-x domain
In Figure A3.1, a solution is supplied at point 1 and leaves at point 2 while vapor is generated
at point 4. This process takes place at the pressure of dew temperature T3. It is assumed that
the solution is saturated.
To begin with, recall from Ch.4 that the molar enthalpy H of a saturated electrolyte solution
at molar concentration x and temperature T can be expressed in relation to that of the
reference solution at xo and T as follows.
⎛ x⎞ ⎛ x⎞ ⎡ m 1 ⎛ ∂φ ⎞ ⎤
H = ⎜ ⎟ H o + ⎜ 1 − ⎟ H H 2O − xυ RT 2 ⎢ ∫ ⎜ ⎟dm ⎥ (A3.1)
⎝ xo ⎠ ⎝ xo ⎠ ⎢⎣ mo m ⎝ ∂T ⎠ ⎥⎦T
In order to use Equation (A3.1) directly, the desorption heat Q& is expressed on molar basis as
The goal is to transform Equation (A3.2) into a simple explicit equation of independent
variables.
Dividing Equation (A3.2) with the rate of vapor generation M& v to get the desorption heat per
unit vapor generation gives
q = (α − 1) ( H 2 − H1 ) + ( H 4v − H1 ) (A3.3)
where ‘q’ and ‘α’ are the desorption heat and the inlet solution flow rate per unit vapor
generation. The flow rate α has the following relation with the bulk concentrations from the
total- and LiBr mass balances between point 1 and 2.
M& x
α ≡ & 1v = 2
(A3.4)
M x2 − x1
For the isotherm T2 in Figure A3.1, letting x=x2 and xo= x1 in Equation (A3.1) gives
⎛ x2 ⎞ ⎛ x2 − x1 ⎞ ⎡ m2 1 ⎛ ∂φ ⎞ ⎤
H 2 = ⎜ ⎟ H1' − ⎜ ⎟ H 3' − x2υ RT2 ⎢ ∫ ⎜
2
⎟dm ⎥ (A3.5)
⎝ x1 ⎠ ⎝ x1 ⎠ ⎣⎢ m1 m ⎝ ∂T ⎠ ⎦⎥T =T2
Inserting Equation (A3.4) and Equation (A3.5) into Equation (A3.3) and rearranging it gives
⎛ x2 ⎞ ⎛ x1 x2 ⎞ ⎡ m2 1 ⎛ ∂φ ⎞ ⎤
q& = ⎜ (
⎟ 1'H − H 1 ) + ( 4 3' ) ⎜ x − x ⎟ 2 ⎢ ∫ m ⎜⎝ ∂T ⎟⎠dm⎥
H v
− H − υ RT 2
(A3.6)
⎝ x2 − x1 ⎠ ⎝ 2 1⎠ ⎢⎣ m1 ⎥⎦T =T2
The enthalpy difference H1'-H1 of the 1st term on the right-hand side of Equation (A3.6) can
be approximated as
⎛ xx ⎞ ⎡ m2 1 ⎛ ∂φ ⎞ ⎤
q& = α C lp1 (T2 − T1 ) + ( H 4v − H 3' ) − ⎜ 1 2 ⎟υ RT22 ⎢ ∫ ⎜ ⎟dm ⎥ (A3.8)
x − x
⎝ 2 1⎠ ⎣⎢ m1 m ⎝ ∂T ⎠ ⎦⎥T =T
2
Now, the last term on the right-hand side of Equation (A3.8) should be simplified. Replacing
φ with Equation (A2.1) and recalling m=x/(1-x)/MH2O, the last term can be rewritten as
−1
⎛ x1 x2 ⎞ ⎡ m2 1 ⎛ ∂φ ⎞ ⎤ ⎛ 1 1 ⎞ ⎡ m2 ∂ ⎛ p* ⎞ ⎛ 1 ⎞ ⎤
⎟υ RT2 ⎢ ∫ ⎜ =⎜ − ⎟ R⎢∫
2
⎜ ⎟dm ⎥ ⎜ ln ⎟d ⎜ ⎟ ⎥ (A3.9)
⎝ x2 − x1 ⎠ ⎣⎢ m1 m ⎝ ∂T ⎠ ⎦⎥T =T2 ⎝ m1 m2 ⎠ ⎣⎢ m1 ∂1/ T ⎝ p ⎠ ⎝ m ⎠ ⎥⎦T =T2
And the gradient inside the integral on the right-hand side of Equation (A3.9) can be
rewritten as
∂ ⎡ ⎛ p* ⎞ ⎤ ⎛ ∂ ln p* ⎞ ⎡ ⎛ ∂1/ T dew ⎞ ⎤
⎢ln ⎜ ⎟ ⎥ = ⎜ ⎟ ⎢1 − ⎜ ⎟⎥ (A3.11)
∂1/ T ⎣ ⎝ p ⎠ ⎦ ⎝ ∂1/ T ⎠ ⎣ ⎝ ∂1/ T ⎠ ⎦
⎛ x1 x2 ⎞ ⎡ m2 1 ⎛ ∂φ ⎞ ⎤
⎟υ RT2 ⎢ ∫ ⎜
2
⎜ ⎟dm ⎥
⎝ x2 − x1 ⎠ ⎢⎣ m1 m ⎝ ∂T ⎠ ⎥⎦T =T2
−1
(A3.12)
⎛ ∂ ln p* ⎞ ⎛ 1 1 ⎞ ⎡ m2 ⎛ ∂1/ T dew ⎞ ⎛ 1 ⎞ ⎤
= ⎜ −R ⎟ ⎜ − ⎟ ⎢ ∫ ⎜1 − ⎟ d ⎜ ⎟⎥
⎝ ∂1/ T ⎠T =T ⎝ m2 m1 ⎠ ⎢⎣ m1 ⎝ ∂1/ T ⎠ ⎝ m ⎠ ⎥⎦
2 T =T2
dp* RT 2 dp* d ln p*
H fg = ΔVT ≈ = −R (A3.13)
dT p dT d1/ T
−1 m
⎛ x1 x2 ⎞ ⎡ x2 1 ⎛ ∂φ ⎞ ⎤ fg ⎛ 1 1 ⎞ ⎡ 2 ⎛ ∂1/ T dew ⎞ ⎛ 1 ⎞ ⎤
⎟υ RT2 ⎢ ∫ ⎜ = H 3' ⎜ − ⎟ ⎢ ∫ ⎜1 −
2
⎜ ⎟dm ⎥ ⎟ d ⎜ ⎟⎥ (A3.14)
⎝ x2 − x1 ⎠ ⎢⎣ x1 m ⎝ ∂T ⎠ ⎥⎦T =T2 ⎝ m2 m1 ⎠ ⎢⎣ m1 ⎝ ∂1/ T ⎠ ⎝ m ⎠ ⎥⎦
T =T2
where the latent heat at the point 3' in Figure A3.1 is defined by Hfg3'≡Hv3'-Hl3'.
Equation (A3.14) can be further simplified if a 1/Tdew-1/T relation is known. For this purpose,
inserting ∂(1/Tdew)/∂(1/T)=a from Equation (A2.4) into Equation (A3.14) gives
⎛ x1 x2 ⎞ ⎡ x2 1 ⎛ ∂φ ⎞ ⎤
⎜ ⎟υ RT2 ⎢ ∫ ⎜
2
⎟dm ⎥ = H 3'fg (1 − a ) (A3.15)
⎝ x2 − x1 ⎠ ⎣⎢ x1 m ⎝ ∂T ⎠ ⎦⎥T =T2
−1 m
⎛ 1 1 ⎞ 2
⎛1⎞
a=⎜ − ⎟
⎝ m2 m1 ⎠
∫ a d ⎜⎝ m ⎟⎠
m1
(A3.16)
Although it was found that Equation (A3.18) is an excellent approximation, the use of
coefficient ā, a highly non-linear function of concentration defined by Equation (A3.16),
makes its application inconvenient.
It was found that the following slightly different equation provides equally good results.
Note that ax = x1 , the gradient a determined at x1, is used instead of ā and Hfg3, the latent heat
at T3, is used instead of Hfg3' in the last term. Also note that T1 replaced T2 in the 2nd term.
Equation (A3.19) has been compared with the values of Equation (A3.3) based on the
enthalpy correlation developed in Appendix A1. The results are illustrated in Figure A2.2 for
the typical working ranges of generator.
1.6
p=13 kPa (Tdew=50oC)
1.2
Error (%)
0.8
0.4
0.6
xw1=0.5
0
0.4 0.5 0.6 0.7
xw2
Figure A3.2 Error of Equation (A3.19)
Error of Equation (A3.19) increases with the concentration change during the process, i.e.
Δx=xw2-xw1. The maximum error would be less than 1.6 %, even when Δx is 0.1.
Multiplying Equation (A3.19) by M& v and expressing it on mass basis gives the desorption
heat Q& finally as
From the thermodynamic point of view, absorption is only the reverse process of desorption
except for the vapor condition in the process. Therefore the results of the section 5.2.2 can be
used for absorption process with a few minor modifications to take account of the different
vapor condition.
Reversing the desorption process, the point 2 in Figure A3.1 now becomes the inlet and the
point 1 becomes outlet of the absorption process. Figure A3.3 shows the absorption process
in a p-T-x domain.
A solution is supplied at point 2 and leaves at point 1 after absorbing the vapor coming from
point 3.
Tdew x=0
3'
T2
M2 T2 x2
x1
2 H1
1' x2
Mv Hv3
Q
3 1 2 1 H2
T3
M1 T1 x1
T3 T1 T2 Tsol
Figure A3.3 Absorption process in p-T-x domain
For the absorption process in Figure A3.3, the absorption heat can be expressed also by
Equation (A3.2) with Hv4 replaced by Hv3. Then, analogous to Equation (A3.20), the
absorption heat Q& for the process in Figure A3.3 can be written as
Error of the Equation (A3.21) is shown in Figure A3.4 for typical working ranges of
absorber, which is comparable to that of generator in Figure A3.2.
3
p=0.87 kPa (Tdew=5oC)
2
Error (%)
0
0.7
xw2=0.6
-1
0.5 0.55 0.6 0.65 0.7
xw1
Figure A3.4 Error of Equation (A3.21)
Appendix B
In most absorption systems, solutions are designed to flow in thin films over various heat and
mass transfer surfaces because it is one of the efficient ways to promote absorption. In this
section, the most commonly used liquid distribution device called dripping-hole tray is
modeled. All the distributors in the chiller under construction were designed in this type.
A dripping-hole tray is an open box or tube with holes at the bottom as shown in Figure B.1.
Tubes with small diameter are sometimes inserted into the holes to guide the flows to fall
onto exact spots. The diameter of tubes and the pitch between them are determined by the
given flow rate and the length to be covered.
The liquid supplied into the tray at a point spreads out sideways and flows through the holes
below. Both the horizontal and vertical velocity U and V in Figure B.1 are functions of liquid
level at the location. The design of a distributor of this type requires the modeling of U and
V.
Liquid supply
V
L/2
Flow through a hole or tube is discontinuous when the level is lower than a certain threshold
value. That is, liquid is dripping when the liquid level is low. This is due to the surface
tension of the liquid. This periodic discontinuous flow is quite common in absorption
systems because distributors are designed to distribute a small amount of liquid evenly along
the long edge of heat exchangers.
Modeling has been carried out separately for the vertical and horizontal velocity. Firstly, the
dripping flow was modeled and the results were used to predict the horizontal flow.
Before advancing to the modeling details, it would be helpful to look into the behavior of a
droplet. An illustration of the flow is given in Figure B.2.
V V V V
(a) W
W
(b)
(c)
(d)
Growing
W
t=0
(e)
Elongating (f)
t1 Thinning
t2 t3
At time t=0, liquid begins to flow through a tube slowly (see Figure B.2a). In this 1st period,
the velocity of the drop W is roughly equal to the growth rate of the droplet’s radius, which is
very small compared to V. As the droplet grows, the pressure within the droplet decreases
(note that pressure inside a droplet is inversely proportional to the droplet diameter) and
consequently the V increases (see Figure B.2b).
As the mass of the drop keep increasing, the surface tension between the edge of the tube
and the droplet is gradually counterbalanced by the weight. The droplet becomes bigger and
W increases rapidly (see Figure B.2c). A liquid column grows between the edge of the tube
and the droplet (see Figure B.2d).
As the droplet velocity W exceeds the flow velocity V, the liquid column begins to thin
because the incoming flow is too small to keep up with the high elongation rate (see Figure
B.2e). Finally the liquid column breaks off right above the droplet and the droplet falls freely
(see Figure B.2f).
Modeling of the flow is done for the three distinctive periods in the process, namely
‘growing’, ‘elongating’ and ‘thinning period’.
Growing period
The liquid begins overcoming the friction along the tube wall and the capillary pressure
inside the curvature of liquid film dangling from the tube tip. The droplet may be assumed to
be a sphere and so its velocity the time derivative of the sphere’s radius. One-dimensional
approach is adopted for two control volumes in Figure B.3. The control volume CV1 is a
circular column that includes the flow field between the free surface and the end of tube. The
control volume CV2 includes the droplet dangling from the edge of the tube.
r
H z CV1
V(t)
Ht
d
do
v(t,r)
CV2
Governing equations for the control volume CV1 and CV2 are coupled in terms of the mass
continuity and the pressure at the plane where the two control volumes meet, i.e. at the end of
the tube.
A governing equation for the flow in CV1 is derived as follows. A detailed diagram for CV1
is given in Figure B.3.
r
z d
do
vz(t,r,z)
σ sinθo
H
τrz
Pn
P(z) θo
V(t)
d
Ht
m=ρQ
τw
fg
D
An
Pn Mean velocity V =
∫ vdA
A W
V(t)
vz(t,r)
dV
ρ = −∇p + ∇τ + ρ g (B.1)
dt
Assuming the flow is axisymmetric and neglecting the normal stress on z-plane τzz, the
momentum equation in z-direction is:
dv z ∂p ∂τ
ρ = − + zr + ρ g (B.2)
dt ∂z ∂r
Assuming vr≈0 and integrating Equation B.2 over the entire control volume gives
⎡∂ ⎛ ∂v ⎞ ⎤
ρ⎢
⎣ ∂t
∫∫∫ v z d(vol) + ∫∫∫ ⎜ v z z ⎟ d(vol) ⎥
⎝ ∂z ⎠ ⎦
= − pn An − τ rzπ d ( H + H t ) + ρ gAn ( H + H t ) (B.3)
≈ − pn An − τ wπ dH t + ρ gAn ( H + H t )
where τ zr is the average shear stress over the entire control surface in z-direction. Since the
shear stress in the tube is much greater, the shear force term may be approximated as
τ wπ dH t .
∫∫∫ v z d(vol) = ∫ ( ∫∫ v dA ) dz = VA ( H + h)
z n n t (B.4)
where V is the average velocity at the end of the tube, which is only a function of time t. h in
the parenthesis is a function of velocity distribution between the free surface and the inlet of
the tube, which should be smaller than the level H. In an extreme case where vz=V
throughout the whole control volume, h is equal to H, which is the case for a confined flow in
a tube. An accurate h would be only available from a two-dimensional flow analysis. In this
study, a linear profile is assumed for the average velocity along the z-axis and thus h=H/2 is
used.
The second integral on the right-hand side in Equation B.3 may be approximated as follows.
1 ⎛ ∂v z ⎞
2
⎛ ∂v z ⎞ 1 2
∫∫∫ ⎜⎝ vz ∂z ⎟⎠ d(vol) = 2 ∫∫ ⎜⎝ ∫ ∂z dz ⎟⎠ dAn ≈ 2 V An (B.5)
Finally, a governing equation for the velocity at the outlet of the tube is given by
dV V 2 p 8ν L
( L + h) + =− n − V + g (H + Ht ) (B.6)
dt 2 ρ d An
where a laminar wall shear stress is used for τw, i.e. τw =8μV/d. pn in Equation B.6, the
pressure at the end of the tube, has to be determined to solve the equation. This pressure can
be removed using the governing equation for the droplet below.
For CV2 in Figure B.3, assuming uniform state inside the control volume and using the
Reynold’s transport theorem, z-momentum equation is given by
⎡∂ ⎤
ρ⎢
⎣ ∂t
∫∫∫ v z d(vol) + ∫∫ VV ⋅ dA s ⎥ = pn An + mg − σ sin θ 0π d 0
⎦
(B.7)
dm dW
W +m − ρV 2 An = pn An + mg − σ sin θ 0π d 0 (B.8)
dt dt
where m and W are the mass and velocity of the droplet respectively. Knowing dm/dt=ρVAn
and rearranging yields
Q dW p Qg σ sin θπ d o
WV + −V 2 = n + − (B.9)
An dt ρ An ρ An
In order to eliminate pn, combining Equations B.6 and B.9 gives a single differential equation
for V as follows.
dV V 2 32μ L Q dW ⎡ σ sin θπ d o ⎤
( L + h) − + (W + )V + ( − g) + ⎢ − g ( H + H t ) ⎥ = 0. (B.10)
dt 2 ρd 2
An dt ⎣ ρ An ⎦
Since V=(dQ/dt)/An, if the relationship between W and Q is given, Equation B.10 can be
expressed as a function of the single variable Q.
In the growing period, when the droplet is growing in radial direction, its shape may be
approximated spherical. Then the velocity of the droplet W is the time derivative of the
sphere’s radius and the following relationships exist.
dQ
W = ( 36π )
−1/ 3
Q −2 / 3 (B.11)
dt
−1/ 3 ⎡ −2 / 3 d Q 2 −5 / 3 ⎛ dQ ⎞ ⎤
2 2
dW
= ( 36π ) ⎢Q − Q ⎜ ⎟ ⎥ (B.12)
dt ⎣⎢ dt 2 3 ⎝ dt ⎠ ⎦⎥
Once Equation B.13 is solved, pn can be calculated back from either Equation B.6 or B.9.
Initially when V is zero, pn has a maximum value, the static hydraulic pressure ρg(H+Ht). As
the V accelerates, the pn decreases and eventually reaches a minimum value, which is
assumed to be atmospheric pressure patm. This is the end of the growing period.
Elongating period
Since pn reached a constant value, Equation B.6 is no longer coupled with Equation B.9. That
is, V is independent from the behavior of the droplet below. Equation B.6 is solved for V and
then Equation B.9 can be solved for W. Shape of the droplet may also be approximated as
sphere in the early part of this period but, as W accelerates, it soon begins to elongate and
looks like Figure B.4c in the last part of the period.
Since W is still smaller than V in this period, the flow through the tube actually pushes the
droplet downwards to grow and accelerate. In the early part of this period, the droplet grows
in radial direction. The tip of the tube is still within the radius of the sphere (see Figure B.4a).
As W increases, the tip of the tube moves out of the spherical boundary and liquid column
grows (see Figure B.4b). The width of liquid column is assumed to be equal to the tube
diameter and constant. Since the relative velocity (W-V) is negative, the droplet still
continues to grow. Finally when W is equal to V, the elongating period is over and the droplet
ceases to grow (see Figure B.4c).
V V V
(a)
r
(b) dr/dt=0
W
(c)
Thinning period
The droplet further accelerates. Since the supply speed of the liquid (V) is less than the
stretching speed (W) of the liquid column, the column begins to thin over the whole length
(see Figure 3.34). It eventually breaks off right above the droplet where the velocity is the
largest. The volume of the droplet is assumed equal to that of the droplet at the end of
elongating period, namely Qdrop.
r0 rl
θl
V Qdrop W
The liquid column may be approximated as a tapered circular column. Then mass balance
equation over the liquid column is given as follows.
π πl drl
(rl 2 + rl r0 + r02 )W + (2rl + r0 ) = π r02V (B.14)
3 3 dt
The momentum equation for W has now only gravity and capillary force terms.
dW 2π rl lσ
mcr = mcr g − σ (2π rl ) sin θ l = mcr g − (B.15)
dt l 2 + (r0 − rl ) 2
where mdrop=ρQdrop. Equation B.14 and B.15 are coupled by rl in the capillary force term in
Equation B.15. rl is assumed to be equal to the hole diameter at the beginning of the thinning
period.
For the growing period, Equation B.13 is solved first for Q. And then pn and W are solved
using Equations B.6 and B.11 respectively. After pn is reduced to atmospheric pressure,
Equations B.6 and B.8 are no longer coupled and thus solved separately for the elongating
period. For the thinning period, Equations B.8, B.9 and B.10 are solved until rl becomes zero,
which is when the droplet breaks off.
An example of the solution is shown in Figure B.6 for water at 18°C through a tube of
φ=1mm and Ht=2mm with the water level H=60mm.
300 1.2 0.06
Pn V
200
0.8 0.04
Velocity(m/sec)
Pressure(Pa)
mdrop
mdrop(gram)
100
0.4 0.02
0 0
0 40 80 120 160
Time(msec)
Figure B.6 Example of solution (T=18°C, H=60 mm, Ht=2 mm, d=1 mm)
At t=0, there is no flow through the tube (V=0) and pn is the static hydraulic head. V increases
with time and pn decreases to atmospheric pressure at 41msec. In the next period, the velocity
of droplet W keeps increasing and eventually becomes equal to V at 167 ms. Liquid column
above the droplet begins to contract and finally breaks off at 178 ms.
Because the velocity V is time-dependent, it is not convenient to use for distributor design.
For that purpose, a time-averaged velocity is more desirable. The average velocity is defined
as follows.
1 t =τ
Vavg =
τ ∫
0
Vdt (B.16)
The time-averaged velocities are calculated for different liquid level H and tube diameters in
Figure B.7 for example. All of the curves show similar trends against the liquid level. It was
found that the results can be excellently represented by
1 ⎛ H − C3 ⎞
H = C1 exp(C2 Vavg ) + C3 or Vavg = ln ⎜ ⎟ (B.17)
C2 ⎝ C1 ⎠
MPE, d=1mm
LPE, d=1mm
GEN, d=1.2mm
LPA, d=1.2mm
60 MPA, d=1.2mm
H(mm)
40
20
Horizontal flow
Using the results of the dripping flow model, the horizontal flow in the dripping tray is
modeled in this section.
Liquid is supplied to an open box tray of a length 2L at the mid point. The box has a single
line of small holes at the bottom, evenly spaced with a pitch λ. A half of the tray is shown in
Figure B.8. The liquid is supplied at x=0. The liquid would flow to the right because the
liquid level has a negative gradient as some liquid falls through the holes at the bottom.
In order to simplify the model, the bottom of the tray is assumed as a porous plate and the
profile of the liquid level is assumed linear with x. These assumptions may be reasonable for
small λ and L.
L
H=f·x+H0, f=(HL-H0)/L
H0 u(x,H)
U0 u(x,z) H(x) CV3 v
y HL
z x Porous plate
v(x) u(x,H)=-v (dz/dx)-1=-v/f
V
λ
Figure B.8 Control volume for horizontal flow in a tray
1 ⎡ 1 ⎤
UH − Q0 + ⎢⎣( H − c)v − ( H 0 − c)v 0 + b ( H 0 − H ) ⎥⎦ = 0 (B.19)
f
⎡ ∂u ∂u ⎤ ⎡ ∂u 2 ∂vu ⎛ ∂u ∂v ⎞ ⎤ ∂P
ρ ⎢u +v ⎥= ρ⎢ + − u ⎜ + ⎟⎥ = − + (∇τ ) x (B.20)
⎣ ∂x ∂z ⎦ ⎣ ∂x ∂z ⎝ ∂x ∂z ⎠ ⎦ ∂x
L 1 L
−U 02 BH 0 + B ∫ vu (x,H) dx = − BWgf ( H 0 + H L ) L − ∫ τ w ( B + 2 H )dx (B.21)
0 2 0
where B is the width of the tray. u(x,H) in the second term on the left is the horizontal
velocity component at the liquid surface and is determined from the mass continuity over an
infinitesimal control volume that includes the surface as in Figure B.8.
L 1 L
∫0
vu (x,H) dx = −
f ∫ 0
v 2 dx
(B.22)
−a
= 2 2 {[b 2 v 2L − 2bv L + 2]( H L − c) − [b 2 v 02 − 2bv 0 + 2]( H 0 − c)}
f b
The last term in Equation B.22 may be approximated as follows using laminar wall stress
correlation, τw =8μU/dh and Equation B.19.
8μ
L
8μ ⎡
L
B ⎤ 8μ ( B + 2 H ) ⎡ H L ⎤
∫
dh 0
U ( B + 2 H ) dx = ⎢ ∫
fd h ⎣ 0
UH (
H
+ 2) dH ⎥
⎦
≈
fd h H
⎢ ∫ UHdH ⎥
⎢⎣ H 0 ⎥⎦
⎧ 1⎡ H ⎤ ⎫
⎪Q ( H − H ) + ⎢( H − c)v − 0 ⎥ ( H − H ) ⎪ (B.23)
⎪ 0 L 0 f ⎢⎣ 0 0 b ⎥ L
⎦
0 ⎪
⎪ ⎪
8μ [ B + ( H L + H 0 )] ⎪ ( H L − c) ⎡ 1 ( H + 3c) ⎤ ( H − c) ⎡ 1 ( H + 3c) ⎤ ⎪
= L 0 0
⎨− ⎢ ( H L + c)vL − ⎥+ ⎢ ( H + c)v −
0 0
⎥⎬
fd h ( H L + H 0 ) / 2 ⎪ f ⎣⎢ 2 4b ⎦⎥ f ⎣⎢ 2 4b ⎦⎥ ⎪
⎪ ⎪
⎪+c ⎡( H − c)v − ( H − c)v + 1 ( H − H ) ⎤ + 1 ( H − H )2 ⎪
⎪ ⎢⎣ L L 0 0 b 0 L ⎥⎦ bf L 0 ⎪
⎩⎪ ⎭⎪
Then finally, by inserting Equation B.22 and B.23 into Equation B.21, the momentum
equation is complete. For purpose of design, the primary variables of interest are the length
of a tray L and velocity distribution. By choosing an appropriate HL, H0 and L can be
determined. HL should be chosen so that the ratio of velocities at both ends, i.e.V0/VL, should
be minimal, which means an even distribution along the length L.
Some results are illustrated in Figure B.9 for 54.2% and 37°C solution at 166 kg/hr. Hole
pitch λ and hole diameter d have been set to15mm and 1.2mm respectively.
0.5
L=465
0.4
Vavg(m/s)
585
0.3
740
0.2
λ =15mm,dhole=1.2mm
0.1
0 200 400 600 800
x(mm)
In Figure B.9, the solution is supplied at L=0. The time-averaged velocity Vavg is calculated
for three trays of different length. For all cases, Vavg decreases as x increases. But it is clearly
seen that the velocity profile is more even for shorter dripping tray. For the 1480mm-long
tray, in Figure B.9, V0/VL is about 1.5 while that of the 0.93m-long tray is 1.05.
Appendix C
Empirical correlations
1/ 3 0.14
⎛ D⎞ ⎛ μ ⎞
Nu = 1.86 ( Re d Pr )
1/ 3
⎜ ⎟ ⎜ ⎟ (C.2)
⎝ l ⎠ ⎝ μw ⎠
1/ 4
⎡ ρ ( ρ − ρ v ) gh fg kl3 ⎤
hT = 0.725 ⎢ ⎥ (C.3)
⎣⎢ μl (nD )(T − Tw ) ⎦⎥
where,
Re Pr ⎡ ⎛ Ga1/ 3 Num ⎞ ⎤ ⎛ s ⎞
m
2 Ga1/ 3 Γ
where t = 2.5871 , Re = and ξ = 0.2623 − 2.95t by Rogers[1981].
π (2 / 3) Re Pr
4/3 4/3
μ
Γ
where Re = by Wilke[1962].
μ