Anda di halaman 1dari 14

Energy 25 (2000) 233–246

www.elsevier.com/locate/energy

A combined power/cooling cycle


Feng Xu a, D. Yogi Goswami a,*
, Sunil S. Bhagwat b

a
Solar Energy and Energy Conversion Laboratory, Department of Mechanical Engineering, P0 Box 116300,
University of Florida, Gainesville, FL 326116300, USA
b
Chemical Engineering Division, Department of Chemical Technology, University of Mumbai, Mumbai 400019,
India

Abstract

A combined thermal power and cooling cycle is proposed. The proposed cycle combines a Rankine
cycle and an absorption refrigeration cycle. It can provide power output as well as refrigeration with power
generation as a primary goal. The new cycle uses very high concentration ammonia vapor in the turbine
which can be expanded to a very low temperature in the turbine without condensation. This cycle uses an
absorption condensation process instead of the conventional condensation process. A parametric analysis
of the proposed cycle is presented in this paper.  2000 Elsevier Science Ltd. All rights reserved.

1. Background

Combined cycles have been proposed in recent years as alternative power cycles for improving
overall energy conversion efficiency. Another recent improvement in thermal power cycles is
based on using mixed working fluids. Using a multicomponent working fluid and/or multipressure
boiling, one can reduce the heat transfer-related irreversibilities and therefore improve resource
effectiveness. Kalina [1] proposed the use of ammonia-water mixtures as the working fluids in
the bottoming cycle of a combined cycle power plant. Although Kalina [1–4] is recognized for
introducing the multi-component working fluid power cycle and for bringing it to its current state
[1–4], Maloney and Robertson [5] studied an absorption-type power cycle using a mixture of
ammonia and water as a working fluid in the early fifties.
A comparison of the multi-component cycle to the Rankine cycle by EI-Sayed and Tribus [6]
shows a 10–20% improvement in thermal efficiency [5]. Since that time, a number of investigators
have studied the use of ammonia-water mixtures in power cycle applications. Marston [7] conduc-
ted a detailed discussion of multi-component cycle behavior. Marston [7] found that the tempera-

* Corresponding author. Fax: +1-352-392-1071.


E-mail address: solar@cimar.me.ufl.edu (D. Yogi Goswami)

0360-5442/00/$ - see front matter  2000 Elsevier Science Ltd. All rights reserved.
PII: S 0 3 6 0 - 5 4 4 2 ( 9 9 ) 0 0 0 7 1 - 7
234 F. Xu et al. / Energy 25 (2000) 233–246

ture at the separator and composition at the turbine inlet are the key parameters for optimizing
the Kalina cycle. Ibrahim and Klein [8] and Park and Sonntag [9] also analyzed the Kalina cycle.
Their studies show the advantages of the Kalina cycle over the conventional Rankine cycle under
certain conditions. Ibrahim and Klein [8] concluded that the Kalina cycle has an advantage over
the conventional Rankine cycle only when heat exchanger NTU is greater than 5.
Since the Kalina cycle uses the conventional condensation process by exchanging heat with
the environment, it puts a constraint on the lowest temperature of the working fluid exiting the
turbine. This constraint can be relaxed if an absorption condensation process is employed.
Rogdakis and Antonopoulos [10] proposed a triple stage power cycle which is similar to the
Kalina cycle. However, they replaced the distillation condensation of the Kalina cycle with an
absorption condensation process. Kouremenos et al. [11] analyzed this absorption type of power
cycle as a bottoming cycle with a gas turbine topping cycle. Since this cycle still uses ammonia-
water vapor mixtures going through the turbine, the exit temperature must be relatively high in
order to avoid condensation in the turbine.
This study analyzes a new cycle as proposed by Goswami [12,13] that retains the advantages
of the Kalina cycle but removes the constraints of both the Kalina cycle and the Rogdakis and
Antonopoulos cycle as identified above.

2. The proposed cycle

The proposed novel cycle combines two thermodynamic cycles, the Rankine cycle and the
ammonia-absorption refrigeration cycle. This cycle uses ammonia-water mixtures as a working
fluid, which reduces the heat transfer irreversibilities, especially for low temperature finite heat
sources such as heat from solar collectors and geothermal heat. The cycle uses a very high concen-
tration ammonia in the turbine, which can be expanded to a very low temperature without conden-
sation. The very low temperature ammonia provides refrigeration. The low temperature ammonia
is condensed by an absorption condensation process. The net effects are the production of both
power and refrigeration, and a reduction of the effective sink temperature. As shown in Fig. 1,
an ammonia-water mixture (state 1) is pumped to a high pressure (state 2). The mixture is heated
to boil off ammonia (state 5), and the vapor is enriched in ammonia by condensing a part of the
vapor in a condenser/rectifier (state 6). The condensate is richer in water and returned to the
boiler. Ammonia is superheated after the condenser/rectifier to raise its temperature (state 7). The
superheated vapor, which is almost pure ammonia, can be expanded in a turbine to exit at a very
low temperature (state 8). After expansion through the turbine to generate power, the ammonia
is brought to the absorption part of the cycle. Low temperature ammonia first provides cooling
in the cooler (state 9). It is then absorbed by the weak solution from the boiler in an absorber,
to form the basic ammonia-water liquid solution to complete the cycle (state 1).

3. Thermodynamic analysis of the proposed cycle

This section gives a thermodynamic analysis of this novel cycle under idealized conditions,
neglecting the irreversibilities associated with the heat transfer and expansion processes. Although
F. Xu et al. / Energy 25 (2000) 233–246 235

Fig. 1. A modified ammonia-based combined power/cooling cycle.

Table 1
Typical operating conditions

State T (K) p (bar) h (kj/kg) S (kl/kg K) X Flowrate


m/m1

1 280.0 2.0 ⫺214.1 ⫺0.1060 0.5300 1.0000


2 280.0 30.0 ⫺211.4 ⫺0.1083 0.5300 1.0000
3 378.1 30.0 246.3 1.2907 0.5300 1.0000
4 400.0 30.0 1547.2 4.6102 0.9432 0.2363
5 360.0 30.0 205.8 1.1185 0.6763 0.0366
6 360.0 30.0 1373.2 4.1520 0.9921 0.1997
7 410.0 30.0 1529.7 4.5556 0.9921 0.1997
8 257.0 2.0 1148.9 4.5558 0.9921 0.1997
9 280.0 2.0 1278.7 5.0461 0.9921 0.1997
10 400.0 30.0 348.2 1.5544 0.4147 0.8003
11 300.0 30.0 ⫺119.0 0.2125 0.4147 0.8003
12 300.0 2.0 ⫺104.5 0.2718 0.4147 0.8003
236 F. Xu et al. / Energy 25 (2000) 233–246

Table 2
Results from Table 1 state conditionsa

Boiler heat input 390.4


Superheat input 31.3
Condenser heat rejection ⫺83.8
Absorber heat rejection ⫺358.8
Refrigeration output 25.9
Turbine work output 76.0
Turbine liquid fraction 0.0692
Turbine vapor fraction 0.9308
Pump work, input 2.7
Total heat input 421.6
Net power and refrigeration output 99.23
Thermal efficiency 23.54%
Carnot efficiency (between 410 K and 280 K) 31.7%
a
All energy units are kW/kg basic solution.

Fig. 2. Effect of turbine inlet pressure on vapor flow rate.

these assumptions limit the usefulness of the analysis, the results show the potential of the pro-
posed cycle for using low temperature sensible (finite capacity) heat sources. The idealized cycle
analysis does provide the analytical maximum limits for the real cycle. Using the thermodynamic
properties data as described in Xu and Goswami [14], the important components of the system
F. Xu et al. / Energy 25 (2000) 233–246 237

Fig. 3. Effect of turbine inlet pressure on turbine work output.

were modeled, and a parametric thermodynamic analysis was conducted with the following con-
ditions:
1. Boiler temperature: 400 K
2. Turbine inlet temperature: 410–500 K
3. Turbine inlet pressure: 18–32 bar
4. Ammonia concentration: 0.20–0.55 by mass in the basic solution

The thermodynamic state conditions of the proposed combined cycle were evaluated assuming
an idealized cycle (that is, irreversibilities associated with real apparatus were neglected).
A computer program for ammonia-water mixture thermodynamic properties was developed for
the cycle analysis. The program uses Gibbs free energy method and has shown good agreement
of the calculated properties with the published literature data. The following assumptions were
made for the cycle analysis:

1. At point 1, the basic solution of ammonia-water mixture is a saturated liquid at low pressure.
For the present analysis, the temperature is set at 280 K to keep the fluid in a liquid state at
the low pressure in the cycle.
2. At point 2, the saturated liquid is pumped to the high pressure of the cycle.
3. The mixture passes through a preheat heat exchanger and the temperature is raised to about
350 K, assuming that the boiler temperature is 400 K.
4. The mixture enters the boiler where it is heated to 400 K. The NH3/H2O mixture evaporates
238 F. Xu et al. / Energy 25 (2000) 233–246

with a higher concentration of NH3. Therefore, at point 4, the ammonia mass concentration is
over 0.90. At point 10, weak aqua returns to the absorber via a heat exchanger.
5. Because we need vapor at a very high ammonia concentration, a part of the water vapor is
condensed in the condenser/rectifier. The condenser/rectifier temperature is set at 360 K.
6. The mixture is superheated before it enters the turbine. Superheater temperature is set at 410
K or higher.
7. The superheated vapor expands in the turbine and transfers work to a generator. The mixture
can be expanded to a very low temperature and still maintain a vapor state.

4. Basic equations

Boiler heat transfer:


qboiler⫽m4h4⫹m10h10⫺m3h3⫺m5h5 (1)

Condenser heat transfer:


qcond⫽m5h5⫹m6h6⫹m3⬘h3⬘⫺m4h4⫺m3⬘h2 (2)

Fig. 4. Effect of turbine inlet pressure on cooling capacity.


F. Xu et al. / Energy 25 (2000) 233–246 239

Fig. 5. Effect of turbine inlet pressure on thermal efficiency.

Superheat input:
qsuperheater⫽m6(h7⫺h6) (3)

Absorber heat rejection:


qabsorber⫽m1h1⫺m12h12⫺m9h9 (4)

Cooling capacity:
qcool⫽m8(h9⫺h8) (5)

Net work output:


wnet⫽m7(h7⫺h8)⫺m1(h2⫺h1) (6)

Thermal efficiency:
wnet+qcool
h⫽ (7)
qsuperheater+qboiler
240 F. Xu et al. / Energy 25 (2000) 233–246

5. Results and discussion

In a simulation of the cycle shown in Tables 1 and 2, with the working fluid entering the
turbine at 410 K and 30 bar pressure and exiting at a 2 bar pressure, a first law efficiency of
23.54% is achieved. By contrast, the Carnot cycle efficiency for the same source temperature
(410 K) and a sink temperature of 280 K is 31.7%. The cycle efficiency of a conventional Steam
Rankine cycle between the same source and sink temperatures will be much lower than both the
Carnot cycle and the new proposed cycle. In addition, a system designed to produce 2 MW of
electrical power using the proposed cycle will produce more than 200 tons of refrigeration. A
parametric analysis of the cycle, described below, shows that the cycle conditions can be optim-
ized within the range of low and medium temperature solar collectors for maximum overall
efficiency. The cycle can also be optimized to provide maximum power, maximum refrigeration,
or a combination of both.

5.1. Effect of turbine inlet pressure

Fig. 2 shows that vapor production goes linearly down as the high pressure in the cycle
increases. The turbine power output also goes almost linearly down as the pressure increases, as
seen in Fig. 3. It is known that the enthalpy drop across the turbine is increased as the pressure
ratio increases. But the enthalpy gains from an increased pressure ratio do not make up for the
drop in the vapor flow rate, hence the turbine work output decreases. Cooling capacity increases

Fig. 6. Effect of boiler temperature on thermal efficiency.


F. Xu et al. / Energy 25 (2000) 233–246 241

Fig. 7. Effect of condenser temperature on cooling capacity.

first as the pressure goes up. Then, due to the decreased vapor flow rate, the cooling capacity
goes down at a pressure of about 28 bar. Fig. 4 shows this trend. The maximum point of cooling
capacity changes with the ammonia mass fraction in the basic solution. It occurs at a higher
turbine inlet pressure for a higher ammonia mass fraction.
Although turbine work output decreases as the pressure increases, the thermal efficiency goes
up first to a maximum and then decreases (Fig. 5). This figure is similar to Fig. 4, which shows
the cycle cooling capacity; however, the conditions for the maximum thermal efficiency do not
coincide with those for the maximum cycle cooling capacity. The maximum cycle thermal
efficiency increases as the ammonia mass fraction increases. However, there is a limit to the
increase in ammonia composition at a given absorber pressure and temperature. The limitation
of the absorber condition on the cycle performance will be discussed in a later section.

5.2. Effect of boiler temperature

The effect of boiler temperature is shown in Fig. 6 at a turbine pressure ratio of 12.5, a con-
denser temperature of 360 K, and a superheater temperature of 410 K. Since the turbine pressure
ratio and inlet temperature are fixed, the enthalpy drop will remain the same regardless of boiler
temperature. Since the vapor flow rate goes up almost linearly as the boiler temperature goes up,
the turbine work output and the cooling capacity increase similarly. The heat input increases
rapidly as the boiler temperature increases. Therefore, the thermal efficiency will reach a limit
even though the turbine power output and cooling capacity increase. To change this limit, the
242 F. Xu et al. / Energy 25 (2000) 233–246

Fig. 8. Effect of condenser temperature on thermal efficiency.

condenser temperature has to increase as the boiler temperature goes up so that more vapor will
remain available to the turbine.

5.3. Effect of condenser/rectifier temperature

The condenser/rectifier temperature controls the vapor ammonia concentration. A lower con-
denser temperature will produce a drier ammonia vapor, which can be allowed to drop to a lower
temperature in the turbine without condensation. The disadvantage is that the vapor flow rate will
also drop.
Fig. 7 shows that the cooling capacity drops as the condenser temperature increases. There is
no cooling available when the condenser temperature is greater than 390 K in the present case.
As the condenser/rectifier temperature goes up, the cooling capacity goes down, but the turbine
work output increases. Fig. 8 shows the change in thermal efficiency with the condenser tempera-
ture, which is due to the combined effect of the changes in the turbine work and the cooling
capacity.

5.4. Effect of superheater temperature

It is expected that turbine work output would increase as the superheat temperature is increased.
The cooling capacity, however, drops as the superheat temperature increases. The reason is that
for a fixed pressure ratio, as the turbine inlet temperature is increased, the exit temperature also
F. Xu et al. / Energy 25 (2000) 233–246 243

increases. There is no cooling capacity available when the superheat temperature is greater than
470 K.
In a conventional Rankine cycle, the thermal efficiency increases as the superheat temperature
increases, since all of the superheat (in the ideal case) is converted to work output. In the novel
cycle, cooling capacity is included in the thermal efficiency. With an increase in the superheat
temperature, the drop in the cooling capacity is steeper than the increase in the turbine work
output. Therefore, the thermal efficiency drops steadily against the superheat temperature (Fig.
9). However, the thermal efficiency starts increasing with the superheat temperature after it reaches
a value (470 K in the present case) where the cycle stops providing any cooling capacity.

5.5. Effect of absorber temperature

The absorber in the proposed cycle takes the place of a condenser in a conventional Rankine
cycle. The temperature of the cooling media limits the absorber temperature. The lower the
absorber temperature, the higher the thermal efficiency will be. In the proposed cycle with the
same ammonia concentration, the turbine exit pressure has to be increased as the absorber tempera-
ture increases in order to condense the ammonia vapor.
Fig. 10 shows the effect of absorber temperature on the thermal efficiency for an ammonia
mass fraction of 0.5. In the present case, no cooling capacity is obtained when the absorber
temperature is 320 K. At 300 K, no cooling capacity is obtained when the turbine inlet pressure
is less than 25 bar.

Fig. 9. Effect of superheat temperature on thermal efficiency.


244 F. Xu et al. / Energy 25 (2000) 233–246

Fig. 10. Effect of absorber temperature on thermal efficiency.

5.6. Other considerations

Fig. 11 shows the effect of turbine inlet pressure on the thermal efficiency for ammonia concen-
trations of 0.20, 0.25, 0.30 and the corresponding turbine exit pressures that would allow conden-
sation in the absorber. Basic solutions with 0.25 and 0.30 ammonia mass fraction are able to
maintain a thermal efficiency of 16–18% in the entire pressure range investigated, while a basic
solution with a 0.20 ammonia mass fraction performs well in a lower pressure range. From Fig.
11, one can determine the best mass fraction in the basic solution under the operating conditions.

6. Conclusions

A combined power/cooling cycle using ammonia-water mixtures as a working fluid is proposed.


The cycle is a combination of Rankine and absorption refrigeration cycles. It will not only produce
power but also provide a certain amount of cooling.
Initial simulation results show that the cycle can achieve high thermal efficiencies for heat
source temperatures around 400 K, which can be obtained easily from geothermal sources, flat
plate and low concentration solar collector, waste heat from other cycles, to name a few. Using
flat plate or low concentration solar thermal collectors in this cycle can reduce the cost of a solar
thermal power plant from $3500/kW at present to less than $2000/kW. If a solar thermal power
plant based on this cycle is combined with a natural gas combustion turbine (especially if natural
F. Xu et al. / Energy 25 (2000) 233–246 245

Fig. 11. Effect of turbine inlet pressure on thermal efficiency for different turbine exit pressure.

gas is used as a backup fuel), which costs around $500 to $700/kW, the combined cost will be
in the range of about $1200 to $1500/kW, which can make solar thermal power cost competitive.
The cycle can be optimized to produce maximum power, maximum refrigeration, or a maximum
overall thermal efficiency. It is recognized that additional studies, especially experimental studies,
are needed to establish the practical usefulness of this cycle.

References

[1] Kalina AI. Combined cycle and waste-heat recovery power systems based on a novel thermodynamic energy cycle
utilizing low-temperature heat for power generation. ASME Paper 1983; 83-JPGC-GT-3.
[2] Kalina AI. Combined cycle system with novel bottoming cycle. Journal of Engineering for Gas Turbines and
Power 1984;106:737–42.
[3] Kalina AI, Tribus M. Advances in Kalina cycle technology (1980–1991): Part I Development of a practical cycle.
Energy for the Transition Age, Proceedings of the Florence World Energy Research Symposium, Firenze, Italy,
1990; 97–110.
[4] Kalina AI, Tribus M, El-Sayed YM. A theoretical approach to the thermophysical properties of two-miscible-
component mixtures for the purpose of power-cycle analysis. ASME Paper 1986; 86-WA/HT-54.
[5] Maloney JD, Robertson RC. Thermodynamic study of ammonia-water heat power cycles. Oak Ridge National
Laboratory Report 1953; CF-53-8-43.
246 F. Xu et al. / Energy 25 (2000) 233–246

[6] El-Sayed YM, Tribus M. A theoretical comparison of the Rankine and Kalina cycles. ASME Special Publication,
1985; AES-1:97–102.
[7] Marston CH. Parametric analysis of the Kalina cycle. Journal of Engineering for Gas Turbines and Power
1990;112:107–16.
[8] Ibrahim OM, Klein SA. Absorption power cycles. Energy (Oxford) 1996;21(1):21–7.
[9] Park YM, Sonntag RE. A preliminary study of the Kalina power cycle in connection with a combined cycle
system. International Journal of Energy Research 1990;14:153–62.
[10] Rogdakis ED, Antonopoulos KA. A high efficiency NH3/H2O absorption power cycle. Heat Recovery Systems
1991;II:263–75.
[11] Koremenos DA, Rogdakis ED, Antonopoulos KA. Cogeneration with combined gas and aqua-ammonia absorption
cycles. In: Krane RJ, editor. Thermodynamics and The Design Analysis, and Improvement of Energy Systems,
vol. 33. New York: ASME AES, 1994:231–8.
[12] Goswami DY. Solar thermal power technology: Present status and ideas for the future. Energy Sources Journal
1996;20:137–45.
[13] Goswami DY, Solar thermal power-status and future directions. Proceedings of the 2nd ASME-ISHMT Heat and
Mass Transfer Conference, Mangalore, India, December 1995.
[14] Xu F, Goswami DY. Thermodynamic properties of ammonia–water mixtures for power-cycle applications. Energy
1999;24:525–36.

Anda mungkin juga menyukai