You are on page 1of 13

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

net/publication/263845643

Comparison of Optimum Operating conditions for a Combined Power and


Cooling Thermodynamic Cycle

Conference Paper · August 2005

CITATIONS READS

3 1,823

3 authors, including:

S.M. Sadrameli
Tarbiat Modares University
92 PUBLICATIONS   2,125 CITATIONS   

SEE PROFILE

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

Thermal Energy Storage View project

DEVELOPMENT OF A RECTENNA FOR ENERGY HARVESTING AND DETECTION APPLICATIONS View project

All content following this page was uploaded by S.M. Sadrameli on 26 October 2014.

The user has requested enhancement of the downloaded file.


APPLIED
Applied Energy 84 (2007) 254–265
ENERGY
www.elsevier.com/locate/apenergy

Optimum operating conditions for a combined


power and cooling thermodynamic cycle
a,*,1 b
S.M. Sadrameli , D.Y. Goswami
a
Department of Chemical Engineering, University of Florida, Gainesville, FL 32611-6300, United States
b
Department of Mechanical and Aerospace Engineering, University of Florida, Gainesville,
FL 32611-6300, United States

Received 2 May 2006; received in revised form 2 August 2006; accepted 5 August 2006

Abstract

The combined production of thermal power and cooling with an ammonia–water based cycle pro-
posed by Goswami is under intensive investigation. In the cycle under consideration, simultaneous
cooling output is produced by expanding an ammonia-rich vapor in an expander to sub-ambient
temperatures and subsequently heating the cool exhaust. When this mechanism for cooling produc-
tion is considered in detail, it is apparent that the cooling comes at some expense to work produc-
tion. To optimize this trade-off, a very specific coefficient-of-performance has been defined. In this
paper, the simulation of the cycle was carried out in the process simulator ASPEN Plus. The opti-
mum operating conditions have been found by using the Equation Oriented mode of the simulator
and some of the results have been compared with the experimental data obtained from the cycle. The
agreement between the two sets proves the accuracy of the optimization results.
Published by Elsevier Ltd.

Keywords: Ammonia cycle; Cooling; Power; Aspen simulation; Optimization

1. Introduction

Multi-component working fluids such as a binary ammonia–water mixture in power


cycles exhibit variable boiling temperatures during the boiling process which makes them
*
Corresponding author.
E-mail addresses: m.sadrameli@gmail.com (S.M. Sadrameli), solar@mae.ufl.edu (D.Y. Goswami).
1
On his leave from Tarbiat Modarres University, Tehran, Iran.

0306-2619/$ - see front matter Published by Elsevier Ltd.


doi:10.1016/j.apenergy.2006.08.003
S.M. Sadrameli, D.Y. Goswami / Applied Energy 84 (2007) 254–265 255

Nomenclature

COP coefficient of performance


Ecool exergy of cooling
Qcool cooling heat transfer
Qh heat input
Rc/w ratio of cooling to work
W net work-output from cycle

Greek
g efficiency
g1st Law cycle’s first-law efficiency definition
gII ref second-law cooling production efficiency

Subscript
work opt optimized for work output only
w/cool dual output conditions

suitable for a sensible heat source. Due to the low temperature-difference between the heat
source and the working fluid, a good thermal match between the source and the working
fluid is allowed, such that less irreversibility occurs during the heat-addition process. This
work has stemmed from commonalities between Kalina-type power cycles and aqua-
ammonia absorption cooling; there is now a small group of proposed configurations [1–
4]. The cited advantages of a combined operation include a reduction in capital equipment
by sharing of components and the possibility of improved resource-utilization compared
to separate power-and-cooling systems [1,2].
A novel ammonia–water binary mixture thermodynamic cycle capable of producing
power and refrigeration has been proposed Goswami [3] and is intended primarily for
power production while simultaneously producing a cooling output. Fig. 1 presents the
schematic diagram of the proposed cycle, where it is shown that the combined cooling out-
put is gained from a heat exchanger following the expander. The expander exhaust is
cooled by expanding the vapor to sub-ambient temperatures. Application of low heat-
source temperatures bellow 200 °C is one of the characteristics of the new cycle in compar-
ison with other dual-output concepts reported in the literature.
Due to the dual outputs of thermal power and cooling, evaluation of the cycle’s perfor-
mance has not been straightforward and requires more consideration. To account for the
quality of the cooling output, which dictates the expander’s exhaust temperature, it was typ-
ically weighted in efficiency definitions [5]. In this work, an analysis of cooling production has
led to a new measure for the effectiveness of cooling production with this cycle. This param-
eter is used as an objective function to optimize the combined power-and-cooling.

2. Background

Ammonia–water mixture as a working fluid for the absorption-cooling cycle applica-


tions, not only has excellent thermo-physical properties, but is an environmentally-friendly
material. It is also the best substitute for the CFCs, for solving global-warming problems.
256 S.M. Sadrameli, D.Y. Goswami / Applied Energy 84 (2007) 254–265

Superheater
Coolant Rectifier
Heat
Source

Separator
Expander

Generator

Heat
Source
Boiler
Cooling Heat Cooled
Exchanger Fluid

Recovery Heat
Exchanger

Throttle

Absorber Coolant

Solution
Pump

Fig. 1. Schematic of the power and cooling cycle.

The thermo-chemical compression system of absorption refrigeration cycles is used as a


heat-driven alternative to the mechanical vapor-compression cycle. It has also been incor-
porated into a power cycle with the ammonia–water working-fluid pair. One of the early
studies of an absorption-based power cycle was performed by Maloney and Robertson [6]
who concluded no significant advantage for the configuration over steam-cycle operation
at the conditions considered. A cycle proposed by Kalina [7] several decades later, is a
well-known absorption power-cycle which is a superior bottoming-cycle option over steam
Rankine cycles. Independent studies have been performed, for example [8,9], that concede
some advantages to the Kalina cycle under certain conditions.
The key benefit of an ammonia–water working fluid is its boiling temperature glide,
which allows for a better thermal match with sensible heat sources and therefore reduces
heat-transfer irreversibilities. However, this temperature glide also exists during the con-
densation process and can limit expansion in the expander, especially at low resource tem-
peratures. Rogdakis and Antonopoulos [10] proposed to take advantage of the chemical
affinity of ammonia–water and replace condensation with absorption–condensation. This
greatly improved performance with low heat-source temperatures because it increased the
amount of expansion that could take place across the expander [10]. This modification
requires that only partial vaporization of the working fluid takes place in the boiler so that
the remaining liquid can be used to absorb vapor in the absorber.
S.M. Sadrameli, D.Y. Goswami / Applied Energy 84 (2007) 254–265 257

Given this configuration, Goswami [3] recognized that, under certain conditions, the
vapor could be expanded to temperatures below those at which absorption–condensation
is taking place. The absorption–condensation process is taking place in the absorber by
using a coolant liquid as shown in Fig. 1. This is an obvious departure from pure working
fluid Rankine-cycle operation, where the limiting expander’s exhaust-temperature is the
vapor condensation temperature, sub-cooling effects aside. The proposal [3] takes advan-
tage of the favorable boiling characteristics of ammonia–water, and capitalizes on the pos-
sible sub-ambient expander exhaust temperatures to form the basis of a combined power-
and-cooling thermodynamic cycle.
Since the original proposal [3], theoretical and experimental investigations have taken
place. Initial investigations were performed theoretically and they focused on identifying
operating trends [11]. Later studies concluded that the cycle could be optimized for work
or cooling outputs and also efficiency. Optimization studies were performed, optimizing
on the basis of the first law, second law, and energy-efficiency definitions [12]. Minimum
cooling temperatures [13], working fluid combinations, and system configurations [14]
were also studied and optimized. An initial experimental study was conducted, which gen-
erally verified the expected boiling and absorption processes [15].

3. Combined cooling/power cycle

Referring to Fig. 1, the relatively strong basic solution of 40 wt% ammonia leaves the
absorber as a saturated liquid at the cycle’s low-pressure and pumped to a high pressure
via the solution pump. Before entering the boiler, the basic solution recovers heat from the
returning weak ammonia liquid solution in the recovery heat-exchanger. As the boiler
operates between the bubble and dew-point temperatures of the mixture at the system
pressure, the basic solution is partially boiled to produce a two-phase mixture: a liquid,
which is relatively weak in ammonia, and a vapor with a high concentration of ammonia.
This two-phase mixture is separated in the separator, and the weak liquid is throttled back
to the absorber. The vapor’s ammonia-concentration is increased by cooling and conden-
sate separation in the rectifier. Heat can be added in the super heater as the vapor proceeds
to the expander. The expander extracts energy from the high-pressure vapor as it is throt-
tled to the system’s low-pressure. The vapor rejoins the weak liquid in the absorber where,
with heat rejection, the basic solution is regenerated.
Several studies on the evaluation of ammonia–water mixture properties have been
reported in the literature. A convenient semi-empirical scheme is used here that combines
the Gibbs free-energy method for mixtures and bubble and dew-point temperature corre-
lations for the phase equilibrium. A comparison between the calculated results and exper-
imental mixture’s properties in the literature has been made by Xu and Goswami [11] shows
a good agreement between two sets of data. By employing absorption–condensation, the
vapor can be expanded to temperatures significantly below the temperature at which
absorption is taking place. Cooling can thus be obtained by sensibly heating the expander’s
exhaust. This behavior is due to the fact that the working fluid is a binary mixture, and at
constant pressure the condensing temperature of an ammonia-rich vapor can be below the
saturation temperature for a lower-concentration liquid. This is best illustrated with a bin-
ary mixture, phase equilibrium diagram, Fig. 2. The low concentration saturated liquid
state with 40 wt% ammonia, approximates the basic solution exiting the absorber, while
the high concentration vapor is typically at the expander’s exhaust conditions.
258 S.M. Sadrameli, D.Y. Goswami / Applied Energy 84 (2007) 254–265

120 Vapor Pressure = 0.203 MPa

100
Two-Phase
80
Temperature [C]

Liquid
60

40 Basic solution
in absorber

20 Expander
exhaust
0

-20
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Ammonia Mass Fraction

Fig. 2. ammonia–water phase equilibrium diagram [18].

4. Cooling

To understand the compromise between work and cooling production with this cycle, a
detailed look at the mechanisms of cooling production are now discussed. In the case of
sensible cooling production, that is no phase change of the refrigerant is occurring, the
exhaust temperature from the expander is the key to determining the amount of cooling
that will be achieved. Also, because the basic solution is only partially vaporized in the
boiler, the quantity of vapor produced is variable and is another factor in determining
the quantity of cooling. The sensitivities of each are now discussed.
Some comments regarding the expander’s exhaust-temperature were discussed in rela-
tion to the binary phase-diagram of Fig. 2. Further insight can be gathered by considering
the entropy of the working fluid at the expander’s exit. Minimization of the exhaust temper-
ature also implies a minimization of the vapor entropy at the expander’s exhaust, assuming
constant exit pressure. From this consideration, an efficient expander is an obvious feature
for low temperatures (low exhaust-vapor entropy of the turbine), but even an ideal device
would only maintain the vapor entropy from inlet to exhaust. Therefore, inlet conditions
should also be considered. For an ammonia–water vapor mixture, entropy decreases with
increasing pressure, increasing ammonia concentration, and decreasing temperature. The
limit of these conditions, while still maintaining vapor, would be saturated pure-ammonia.
Considering these preferred expander’s inlet-conditions, the function of the rectifier, is
immediately apparent. In the rectifier, the vapor’s ammonia concentration is increased by
removing the small amounts of water that have also been vaporized. This is generally per-
formed by taking advantage of the saturation properties of the mixture, which in its sim-
plest form is cooling and condensate separation. The net change to the vapor is an increase
in concentration and a decrease in temperature, accompanied by a minor drop in pressure
and some reduction to the mass flow rate. These effects are mostly to the advantage of the
cooler’s expander exhaust-temperatures.
S.M. Sadrameli, D.Y. Goswami / Applied Energy 84 (2007) 254–265 259

0.9
Vapor Concentration [kg/kg]

0.8 Saturated liquid


conditions
0.7

Complete vaporization
0.6

0.5

0.4

0.3
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Vapor Mass Fraction (kg/kg)

Fig. 3. Relationship between vapor mass fraction (kg/kg) and amount of partial vaporization.

The considerations for exhaust temperatures are now coupled with the mechanisms of
vapor production. The working fluid was partially vaporized in the boiler and separated
by phase in the separator. From initial to complete vaporization, the boiling process pro-
ceeds as indicated in Fig. 3, which is a plot of vapor concentration as a function of the
vapor mass-flow fraction (the ratio of the vapor mass-flow to the basic solution mass-
flow). As shown, with minimal vaporization, the concentration is highest, while at high
amounts of vapor production, the concentration approaches the basic solution concentra-
tion (0.4 ammonia (kg/kg) in this case). From the previous discussion of expander inlet
conditions, high concentrations are preferred, which implies low vaporization rates.
These results imply that, for cooling production the partial-boiling operation should
approximate to a distillation process separating ammonia from the working-fluid mixture.
These are the same general requirements as of an aqua-ammonia absorption cooling cycle;
however, they are in contrast to Rankine-based power production where the production of
vapor, regardless of composition, is a critical consideration.

5. Cycle performance

The fact that this cycle has dual outputs of power and cooling has raised questions
regarding the evaluation of its performance. Central to these is the treatment of the quality
of both the work and cooling outputs. Previous studies in this area are now discussed
along with a description of the evaluation used in this study.

5.1. Prior work

The question of appropriate efficiency expressions for the cycle was examined by Vijay-
araghavan and Goswami [5]. It was noticed that the results obtained from an optimization
260 S.M. Sadrameli, D.Y. Goswami / Applied Energy 84 (2007) 254–265

of the cycle were influenced by the weight given to the cooling output in the objective func-
tion, which was typically an efficiency definition. In its simplest form, a first-law efficiency
could be defined, where the work and cooling outputs of the cycle are added together as
useful outputs,
ðW net þ Qcool Þ
g1st Law ¼ ð1Þ
Qh
However, by not accounting for the quality of the cooling output, Eq. (1) gives an over-
estimate of system’s performance [5]. The quality of the cooling output can be evaluated
by using a COP to find a work equivalent for the cooling produced,
ðW net þ ðEcool =gII ref ÞÞ
g1st Law ¼ ð2Þ
Qh
The Carnot or Lorenz theoretical-efficiency, based on the appropriate temperature lim-
its, could be used for this purpose. However, this would compute the theoretical mini-
mum amount of work needed to produce the cooling, and thus greatly undervalues
the cooling output [5]. A practically-obtainable COP value could be substituted; how-
ever, there is a subjective element to this approach. The authors concede that ultimately
the value of work and cooling, that is the weighting, will be best decided by the end
application [5].

5.2. Effective COP

In this study, a different approach is used to evaluate the effectiveness of combined cool-
ing production. The discussion of cooling-production mechanisms revealed that some
compromises to work production are needed in order to obtain a simultaneous cooling
output. For instance, conditioning of the vapor to accommodate cooling typically reduces
the amount of vapor entering the expander, which reduces the amount of work done. A
new coefficient of performance, specific to this cycle, is proposed based on the idea that
a compromise is needed for the dual outputs of power and cooling. This COP definition
relates the cooling produced with this cycle to the theoretical amount of work production
that was compromised in order to have combined-cooling production. The general formu-
lation is provided as follows:
Cooling produced
COPeffective ¼ ð3Þ
Potential-work lost
This term is referred to as the effective COP since cooling and work are only indirectly re-
lated, in other words, there is no device directly producing cooling with the work that is
not produced. Rather, the effective COP is a measure of the compromise made to accom-
modate a combined cooling output. Just as with conventional COP values, higher values
of effective COP are desired since they maximize the cooling production for the penalty of
less work production. The formulation with cycle parameters is presented as Eq. (4). In
evaluating the lost potential-work, the difference between the work produced with cooling
and the ideal work production is used. The ideal work production, Wwork opt, is the work
output for a work-optimized (based on a first-law efficiency, but giving no value to cool-
ing) power-cooling cycle operating between equivalent heat-source and rejection
conditions.
S.M. Sadrameli, D.Y. Goswami / Applied Energy 84 (2007) 254–265 261

Qcool Rc=w
COPeffective ¼ ¼ g  ð4Þ
W work opt  W w=cool work opt
1
gw=cool

The rightmost expression in Eq. (4) is an equivalent way to formulate the effective COP
and it was used to evaluate the presented results.

6. Modeling

The modeling for this work was based on the connection schematic of Fig. 1 and was
performed using the Equation Oriented (EO) mode in ASPEN Plus, version 12.1 [16]. EO
modeling is a different strategy for solving flow-sheet simulations in Aspen Plus, which is a
very effective way of solving certain kinds of problems, such as process optimization. In
this approach instead of solving each block of simulation in sequence (Sequential Modu-
lar, SM), EO gathers all the model equations together and solves them at the same time.
The PSRK (Predictive Soave–Redlich–Kwong) equation-of-state (EQS) for the property
algorithm within Aspen Plus was used to compute the thermodynamic properties of the
ammonia–water working-fluid. This EQS formulation gave predictions that agreed with
IGT (Institute of Gas Technology) experimentally-measured data [17]. Some of the simu-
lation runs performed before optimization to for the verification of the developed model
have been compared with the experimental data reported by Martin [18]. Averaged con-
ditions for the experimental and simulated runs are as follows:

Expander’s inlet-pressure = 0.0516 MPa


Expander’s exit-pressure = 0.2080 MPa
Vapor concentration = 0.9930 kg/kg
Solution pump flow-rate = 0.0020 kg/s

Noted procedures and constraints for the optimization were as follows:

 The boiling temperature and pressure were specified. The basic solution concentration
was determined by specifying the pump’s inlet-temperature and pressure along with the
assumption of a saturated liquid. The vaporization fraction (vapor quality at the boi-
ler’s exit) was determined from saturation conditions.
 The absorber’s exit-temperature (pump inlet) was assumed to be approximately 5 °C
higher than the assumed ambient temperature of 25 °C.
 Vapor rectification was limited by a specified minimum rectifier exit-temperature of
35 °C.
 The minimum amount of vapor leaving the rectifier was allowed to be 5% of the basic
solution flow-rate.
 The quantity of cooling produced (if any) was calculated as the heat needed to raise the
expander’s exhaust-temperature to the ambient temperature of 25 °C.

The model used for this work has been simplified slightly when compared with the model
used by Vijayaraghavan and Goswami [5]. In that work, the cooling needed in the rectifier
was provided by a diverted portion of the basic solution stream from the absorber, thus pro-
viding some heat recovery. In this work, the rejected heat from the rectifier was judged to be
small compared with the boiler’s input and has been rejected to an external cooling source.
262 S.M. Sadrameli, D.Y. Goswami / Applied Energy 84 (2007) 254–265

7. Results

Fig. 4 shows the ASPEN Plus flow-sheet for the cycle simulation. Rectifier and absorber
were considered as simple separation columns with heating and cooling utilities, to be
identical to the distillation columns in the simulation. The following parameters and
assumptions were used in the simulations: isentropic efficiencies of the turbine, and pump
80%, mechanical and electric efficiency 96%, a minimum temperature-difference for each of
the heat exchangers of 10 °C, ambient temperature 25 °C, ambient pressure 1.0132 bar,
and finally rectifier and separator temperatures of 39 °C and 83.5 °C, respectively. Table
1 shows one of the results for the turbine’s inlet-temperature of 39.1 °C. The comparison
between the simulated and experimental results [18] for the turbine’s inlet-temperatures of
39.1 °C, 40.1 °C, 42.2 °C, and 47.9 °C are shown in Table 2. The agreement between two
sets of the results validates the accuracy of the developed simulation model. An optimiza-
tion of the cycle conditions was performed for a boiler’s exit-temperature of 122 °C and
using the effective COP, Eq. (4), as the objective function. For a fixed value of the pump’s
inlet-pressure, the values of Rc/w (parameter in Eq. (4) have been calculated for different
values of the pump’s outlet-pressure. The calculations were repeated for different pump-
inlet pressures from 2.5 to 5.0 bars until the optimum conditions were reached. The results
are shown in the contour plot of Fig. 5. As can be seen from the figure, the maximum value
of the effective COP was reached at 0.647, with a basic ammonia concentration of 0.554.
The optimum conditions obtained are as follows:

Boiling pressure = 2.3 MPa


Absorber pressure = 0.42 MPa

Fig. 4. Major components in the Aspen Plus [16] process-model for the ammonia cycle.
S.M. Sadrameli, D.Y. Goswami / Applied Energy 84 (2007) 254–265
Table 1
Simulation results by Aspen Plus [16] for one of the operating conditions
1 2 3 4 5 6 7 8 HE-O Pump-O Rec-O Valv-O Valv1-O
Temperature (°C) 31.74 68.67 84.75 83.50 83.50 39.10 22.18 25.00 46.23 31.80 39.10 38.05 46.29
Pressure (bar) 1.81 5.16 4.66 4.66 4.66 4.45 2.08 2.08 4.66 5.16 4.45 4.28 2.08
Vapor frac (kg/kg) 0.00 0.00 0.02 0.00 1.00 1.00 1.00 1.00 0.00 0.00 0.00 0.00 0.00
Mole flow (kmol/h) 0.41 0.41 0.41 0.41 0.00 0.00 0.00 0.00 0.41 0.41 0.00 0.00 0.41
Mass flow (kg/h) 7.31 7.31 7.31 7.25 0.07 0.06 0.06 0.06 7.25 7.31 0.01 0.01 7.25
Ammonia (mol%) 0.24 0.24 0.24 0.23 0.91 0.99 0.99 0.99 0.23 0.24 0.49 0.49 0.23
Water (mol%) 0.76 0.76 0.76 0.77 0.09 0.01 0.01 0.01 0.77 0.76 0.51 0.51 0.77
Ammonia (kmole/h) 0.10 0.10 0.10 0.10 0.00 0.00 0.00 0.00 0.10 0.10 0.00 0.00 0.10
Water (kmole/h) 0.31 0.31 0.31 0.31 0.00 0.00 0.00 0.00 0.31 0.31 0.00 0.00 0.31

263
264 S.M. Sadrameli, D.Y. Goswami / Applied Energy 84 (2007) 254–265

Table 2
Comparison of the simulated and experimental operating-conditions
Operating Turbine Absorber Turbine Boiler Boiler in Separator Separator
Condition inlet (°C) exit (°C) exit (°C) out (°C) (°C) liquid out (°C) vapor out (°C)
Experimental 39.1 31.74 21.95 84.75 68.67 83.18 84.24
Modeling 39.1 31.7 22.2 84.8 68.7 83.5 83.5
Experimental 40.1 31.4 23.4 85.3 68.6 82.5 83.8
Modeling 40.1 31.4 24.5 85.3 68.6 83 83
Experimental 42.2 31.76 24.5 83.7 65.0 81.9 83.2
Modeling 42.1 31.8 24.1 83.7 65.1 82.5 82.5
Experimental 47.9 32.1 26.9 84.31 67.7 82.43 83.5
Modeling 47.9 32.1 27.2 84.3 67.7 83 83

2.4
Effective COP
2.2 0.4
Boiling Pressure [MPa]

2 0.64

1.8 0.60

1.6
0.55

1.4
0.5
Eff COP optimum, 0.647
1.2 [5]
[5]1st 30
optimum, η II ref = 30%
0.40
[5]
[5]1st 50
optimum, η II ref = 50%
1
0.345 0.395 0.445 0.495 0.545
Basic Solution Concentration [kg/kg]

Fig. 5. Operation map showing optimized region and comparison points (COPopt = 0.647 at a concentration of
0.554, expander’s outlet-temperature of 0.1 °C and pump pressure of 2.3 MPa.).

Basic concentration (ammonia wt%) = 0.554


Expander’s inlet-temperature = 69.0 °C
Expander’s exit-temperature = 0.10 °C
Boiler’s vapor fraction = 0.325
Ratio of cooling to work = 0.888

These results were obtained by considering a similar configuration (discussed in the


modeling section) of the power and cooling cycle, simulating with equivalent heat-source
and absorber conditions, but using a combined first-law efficiency definition, Eq. (2), as
the objective function [5].

8. Conclusions

A thermodynamic analysis of a combined cooling-and-power cycle was carried out


using an EO mode of simulation by ASPEN Plus 12.1 [16]. The results show that the con-
S.M. Sadrameli, D.Y. Goswami / Applied Energy 84 (2007) 254–265 265

ditions for the optimum cooling production are at odds with the power production. In this
study, the trade-off of work to accommodate cooling is quantified by defining an effective
COP. The defined parameter has been used as an objective function to optimize the cool-
ing production. It appears that the effective COP optimum is biased towards higher expan-
der exhaust-temperatures but with increased vapor flow. Some of the simulation runs have
been performed before optimization for the verification of the developed model. The
agreement between the simulation and experimental results proves the accuracy of the
developed model. Finally, this analysis is not sufficient to determine the optimum param-
eters for the operating conditions. The optimum conditions have to be tested experimen-
tally for the final verification of the results. Furthermore, it is not clear what conditions
would favor a particular objective function. Exchanging the work-optimized cycle-effi-
ciency in Eq. (4) with an analytic definition, such as Carnot efficiency, would aid in an ana-
lytical comparison of the effective COP with other performance evaluations.

References

[1] Erickson DC, Anand G, Kyung I. Heat-activated dual-function absorption cycle. ASHRAE Trans
2004;110(1):515–24.
[2] Amano Y, Suzuki T, Hashizume T, Akiba M, Tanzawa Y, Usui A. A hybrid power-generation and
refrigeration cycle with ammonia–water mixture, IJPGC2000-15058. In: Proceedings of AD 2000 joint power
generation conference. ASME; 2000.
[3] Goswami DY. Solar thermal power: status of technologies and opportunities for research, heat-and-mass
transfer ’95. In: Proceedings of the 2nd ASME-ISHMT heat-and-mass transfer conference. New Delhi
(India): Tata-McGraw Hill Publishers; 1995. p. 57–60.
[4] Zheng D, Chen B, Qi Y. Thermodynamic analysis of a novel absorption power/cooling combined cycle. In:
Proceedings of the 2002 international sorption heat-pump conference. Shanghai, China; 2002. p. 204–9.
[5] Vijayaraghavan S, Goswami DY. On evaluating efficiency of a combined power-and-cooling cycle. In:
Proceedings of IMECE 2002. New Orleans (LA): American Society of Mechanical Engineers; 2002.
[6] Maloney JD, Robertson RC. Thermodynamic study of ammonia–water heat and power cycles. ORNL
Report CF-53-8-43, Oak Ridge (TN); 1953.
[7] Kalina AI. Combined-cycle system with novel bottoming cycle. ASME J Eng Gas Turb Power
1984;106:737–42.
[8] Marston CH. Parametric analysis of the Kalina cycle. J Eng Gas Turb Power 1990;112:107–16.
[9] Ibrahim OM, Klein SA. Absorption power cycles. Energy 1996;21(1):21–7.
[10] Rogdakis ED, Antonopoulos KA. A high efficiency NH3/H2O absorption power-cycle. Heat Recov Syst
CHP 1991;2:263–75.
[11] Goswami DY, Xu F. Analysis of a new thermodynamic cycle for combined power-and-cooling using low-
and mid-temperature solar collectors. J Solar Energ Eng 1999;121:91–7.
[12] Lu S, Goswami DY. Optimization of a novel combined power/refrigeration thermodynamic cycle. J Solar
Energ Eng 2003;125:212–7.
[13] Lu S, Goswami DY. Theoretical analysis of ammonia-based combined power/refrigeration cycle at low
refrigeration-temperatures. In: Proceedings of SOLAR 2002. Reno (NV): American Society of Mechanical
Engineers; 2002.
[14] Vijayaraghavan S. Thermodynamic studies on alternate binary working-fluid combinations and configura-
tions for a combined power-and-cooling cycle. Ph.D. dissertation, University of Florida, Gainesville (FL);
2003.
[15] Tamm G, Goswami DY. Novel combined power-and-cooling thermodynamic cycle for low-temperature heat
sources, part II: Experimental investigation. J Solar Energ Eng 2003;125:223–9.
[16] Aspen Plus, Version 12.1. Ten Canal Park, Cambridge (MA) 02141: Aspen Technology, Inc; 2004.
[17] Macriss R, Eakin B, Ellington R, Huebler J. Physical and thermodynamic properties of ammonia–water
mixtures. Chicago (IL): Institute of Gas Technology; 1964.
[18] Martin CL. PhD thesis, Mechanical Engineering Department, University of Florida; 2004.

View publication stats

You might also like