Sie sind auf Seite 1von 17

Energy Conversion & Management 40 (1999) 18991915

www.elsevier.com/locate/enconman

Performance modelling of a carbon dioxide removal system for power plants


Umberto Desideri*, Alberto Paolucci
Universita di Perugia, Dipartimento di Ingegneria Industriale, Via G Duranti 1 A/4-06125, Perugia, Italy

Abstract In this paper, a carbon dioxide removal and liquefaction system, which separates carbon dioxide from the ue gases of conventional power plants, was modelled. The system is based on an amine chemical absorption stripping system, followed by a liquefaction unit to treat the removed CO2 for transportation and storage. The eect of the main parameters on the absorption and stripping columns is presented. The main constraints set for the model are a capture eciency of 90% and the use of an aqueous solution with a maximum 30% amine content by weight. The goal of this study is to remove the CO2 with minimum energy requirements for the process when it is integrated in a fossil fuel red power plant. Results of the simulation are compared to experimental and literature data from feasibility studies and existing plants. The power plant to which the removal system is connected is a 320 MW steam power plant with steam reheat and 8 feedwater heaters. Two dierent fossil fuels were considered: coal and natural gas. The eect of the modications necessary to integrate the CO2 removal system in the power plant is also studied. The capital cost of the removal and liquefaction system is estimated, and its inuence on the cost of generated electricity is calculated. # 1999 Elsevier Science Ltd. All rights reserved.
Keywords: Carbon dioxide removal; Chemical absorption; Amines; Power plant

* Corresponding author. Tel./fax: +39 0755852736. E-mail address: desideriu@asme.org (U. Desideri) 0196-8904/99/$ - see front matter # 1999 Elsevier Science Ltd. All rights reserved. PII: S 0 1 9 6 - 8 9 0 4 ( 9 9 ) 0 0 0 7 4 - 6

1900

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

1. Introduction At the Kyoto Conference, many countries in the world recently agreed to reduce emissions of greenhouse gases into the atmosphere or at least to keep them at the current level. Carbon dioxide has been proven to be a greenhouse gas, contributing to the increase of the earth's surface temperature and producing long term climate changes. The EPA has estimated the anthropogenic CO2 emissions to be higher than 6 Gt/yr, half of which are produced by industry and power plants using fossil fuels. Several techniques to remove CO2 from gas mixtures have been studied since 1970, but most of them were applied to produce technical CO2 as process gas, mainly for the food or chemical industry. In the following decade, some of the CO2 capture systems were considered for application in power plants, and a small number of pilot plants were built to prove the technology and verify theoretical calculations [16]. Since most of those systems work better with higher CO2 concentrations than those in the ue gases from fossil fuel red plants, modications to the conventional schemes of power plants were proposed. In particular, closed or semi-closed gas turbines and combined cycles are a promising way of increasing the CO2 concentration in ue gases [79]. Closed and semiclosed gas turbine cycles, however, need a major redesign of power plant components, which is currently too expensive to be justied by the possible market share. Moreover, this is only possible in newly built power plants. The addition of a CO2 removal system to an existing power plant can still be considered the only option for short or medium term interventions to reduce carbon dioxide emissions [10]. The technical feasibility of CO2 removal systems has been proven by several applications in the food, oil and chemical industries, and its application to power plants must only be weighed against the performance reductions and cost increases. Little modications are, in fact, necessary to integrate a CO2 removal system in a conventional steam power plant. Among the alternatives for CO2 capture, chemical absorption with amine aqueous solutions was demonstrated as one of the most mature and less expensive technologies to be applied to power plants [11]. The absorption stripping system is particularly interesting because of its possibility to regenerate the solution continuously, thereby operating in an almost closed cycle. A model for simulation of the absorption and stripping columns, together with a liquefaction unit, has been created by means of ASPEN Plus software.

2. Carbon dioxide removal system The plant for removing carbon dioxide from ue gases has two main elements, which are the absorption and stripping packed columns. This allows a continuous regeneration of the amine solution, which saves considerable amounts of solvent. Amines in the water solution react with CO2 in the absorption column, forming chemical compounds that separate CO2 from the gas mixtures at a higher rate than the natural CO2 absorption in pure water. Dierent amines were applied and studied in the literature [7,1115]. In this paper, the only amine considered was monoethanolamine (MEA), which is a primary amine that produces the

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

1901

carbamate ion when it reacts with CO2. The main advantages in using MEA/water solutions are its high CO2 reactivity, its high limit load (0.5 mole of CO2 per mole MEA) and its low molecular weight. Its main drawback is the stability of the carbamate ion that makes the regeneration more heat demanding. The absorption column was modelled by using ASPEN Plus software. The main assumptions used in the simulation are: . . . . No pollutants in ue gases, such as NOx or SOx, and perfect combustion of the fossil fuel; Adiabatic absorption process; Highest limit of MEA concentration in water at 30% by weight; Henry's law type mass transfer for gas to liquid phase for CO2, O2, N2.

The presence of NOx and SOx must be avoided because they react with amines, preventing regeneration and raising the MEA make-up to cover additional losses. The limit of 30% by wt of MEA in water was set to avoid corrosion, which can be present with richer amine solutions.

Fig. 1. Flow sheet of the CO2 recovery system.

1902

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

Fig. 1 shows a ow sheet of the absorber stripper system. The ue gas entering the absorption column ows up through the vessel, countercurrent to the aqueous solution. The carbon dioxide in the ue gas reacts chemically with the solvent while the puried gas is vented to the atmosphere, and the solvent enriched by CO2 is pumped from the contact tower to a heat exchanger. The rich solvent is preheated in the lean/rich exchanger by the hot lean solution returning from the regenerator/stripper on its way back to the absorber. The rich solvent solution enters the top of the regenerator where it ows down through the vessel countercurrent to the stripping steam generated in the solution reboiler. Steam and solvent vapours move up the regenerator column, condensing as CO2 is liberated and the solvent solution is heated. Uncondensed steam and carbon dioxide leave the top of the regenerator and then enter the reux condenser. The condensate is returned to the system while the carbon dioxide is removed to further processing. The lean solvent solution is pumped from the bottom of the regenerator directly to the lean/rich exchanger. The solvent leaves the lean/ rich exchanger after giving up heat to the rich solution and then enters a cooler, where its temperature is further lowered before being returned to the absorber. The basic simulation has been improved by adding a mixer in order to cover the losses of water and MEA from the gaseous streams leaving the two columns. The main reactions that take place in aqueous systems of amine and CO2 are: 2H2 O , H3 O OH CO2 2H2 O , H3 O HCO 3
2 HCO 3 H2 O , H3 O CO3

2MEA CO2 , MEAH MEACOO MEA H2 O , MEA H3 O MEACOO H2 O , MEA HCO 3

3. Absorption system study The performance of the absorber as a stand alone column was analysed before studying the complete removal system. The eect of ue gas and lean solution temperatures was analysed to determine their inuence on CO2 capture and amine and water evaporation losses. The absorber was simulated as a 7 stage column, where ue gases enter from the 7th stage at the bottom of the vessel, and the lean solution enters at the top in the 1st stage. Two dierent ue gas compositions were assumed in this study: one from the combustion of coal and another from the combustion of natural gas. In the study of the absorber alone, combustion of

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

1903

coal was assumed, giving the following composition of ue gas: 4.5% by mass H2O, 19.5% CO2, 71% N2, 5% O2. The aqueous solution contains 30% MEA by weight, and a liquid solution/gas ratio of 2.0 was calculated as optimal for maximum absorption. The rst step of the analysis consisted in the calculation of the eect of lean solution temperature at the 1st stage inlet on top of the absorber. The lean solution ows back from the stripping column after the amine regeneration process. The lean solution temperature was varied between 30708C, keeping the ue gas temperature unchanged at 658C. The response of the absorber to the MEA+H2O temperature variations at the rst stage is practically linear (Fig. 2) for both the MEA and H2O losses. When the lean solution temperature changes from 30 to 708C, the ue gas temperature in the absorber has a 108C temperature rise, passing from about 64.88C to 74.18C. The rich solution temperature at the vessel bottom is practically unchanged. The absorbed CO2 varies from 17.218.1% by mass, but the water and MEA evaporation losses increase, owing to the higher temperature of gas at the top of the vessel. The water losses double in the studied temperature range of the lean solution, whereas the MEA losses are three times as much. It is then very important to keep the lean solution temperature as low as possible, with two benets: (1) to reduce MEA and water make-up and (2) to increase CO2 capture eciency. A lean solution temperature of 408C is a good compromise between the above benets, cooling duty and cooling tower size. Having assumed a cooling water temperature of 208C, lowering the lean solution temperature under 408C would substantially increase the cooling systems heat exchanger size, without signicant benets to the absorption process. The eect of ue gas temperature was studied in the range 50858C (Fig. 3). Following the considerations described above, the lean solution temperature was kept constant at 408C. The response of the main system parameters is still linear, but with some dierences from the situation described above. The vented gas temperature on top of the absorber is practically constant (Fig. 3a), and the rich solution at the bottom also experiences small temperature

Fig. 2. Inuence of lean solution temperature on MEA and water losses.

1904

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

Fig. 3. Eect of rich solution temperature on the performance of the absorber.

variations (about 48C). MEA and water evaporation losses increase with ue gas temperature, but to a lesser extent than with the lean solution temperature (Fig. 3bc). The vented CO2 ow rate changes from 13.8 (17%) to 15.0 t/h (18.4%) (Fig. 3 d). Therefore, the ue gas temperature at the exit from the steam generator was assumed at 408C for the rest of the paper. A temperature increase of 258C was assumed in the fan, which is necessary to bring the ue gas pressure to 1.2 bar. The temperature of the ue gas entering the absorber is then 658C, a level which was kept constant for the rest of the paper. Changes of pressure in the absorber were not considered because from a preliminary study, it has been shown that the increase of absorption eciency with a 2.4 bar absorber pressure is not justied by the higher power required by the fan.

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

1905

4. Stripping system study The MEA aqueous solution, enriched by CO2 and coming out of the absorber, ows into a shell and tube heat exchanger (HEAT-EX) where it is heated by heat transfer from the lean solution coming from the stripper. The hot uid (lean solution) ows in the tube side, and the cold uid (rich solution) ows on the shell side. This heat exchanger reduces the cooling duty on the lean solution by internal regeneration of heat. The stripper column was modelled with 11 theoretical stages, where the condenser represents the rst stage and a kettle reboiler represents the 11th stage. The heat required by the rich solution regeneration process is provided to the kettle reboiler by steam extracted from the power plant connected to the CO2 removal plant. This is the main source of performance reduction of the power plant. The heat source in other applications can be a conventional industrial steam generator. The condenser temperature on top of the stripper was set at 908C. This temperature is a trade-o between the reduction of the water ow rate exiting with CO2 towards the liquefaction unit and the temperature of the uid returning to the column, which should be high enough for good regeneration eciency. Connected to the stripping temperature is the stripping pressure because an increase of the latter raises the temperature in the reboiler and, subsequently, in the whole column. Fig. 4 shows the recovered CO2 and the make-up water for dierent stripper pressures. The curves were calculated by keeping the reboiler and condenser heat duties unchanged, when treating 272 t/h of ue gas with a CO2 mole fraction of 8.55%. Increasing stripper pressure (IN-STRIP ow) improves regeneration eciency and reduces CO2 compression work in the subsequent liquefaction process. At the same time, higher stripper pressures increase the boiling point of the rich solution, and the reboiler has to be fed with steam extraction from the power plant at a higher pressure. Changing the liquid solution to ue gas ratio (L/G) can further reduce the heat duty of the stripping process. Table 1 shows a comparison of two L/G ratios used for treating ue gas

Fig. 4. Inuence of stripper pressure on recovered CO2 and water make-up.

1906

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

Table 1 Energy consumption of a CO2 separation system where ue gases from the combustion natural gas and coal are treated Flue gases from natural gas combustion: CO2 content=8.55% by mole L/G ratio 2.00 Reboiler heat duty (MW) 39 Heat duty per ton recovered CO2 GJ=tCO2 4.22 Flue gases from coal combustion: CO2 content=13.1% by mole L/G ratio 2.48 Reboiler heat duty (MW) 69 Heat duty per ton recovered CO2 GJ=tCO2 4.68 2.286 36 3.88 2.74 58.5 3.95

from natural gas and coal combustion. The table also shows that a higher L/G ratio reduces the heat duty per ton of recovered CO2, while the increase in power required by pumps is negligible in comparison with that required by the gas compressors located before the absorber. 5. Comparisons with literature data The CO2 removal model was validated against data available in the literature from case studies and pilot plant experimental results. Most of the published sources do not provide all the data necessary for a complete comparison of the results obtained in the present paper. Some of the papers do not report the exact composition of the solvent, and none indicates the liquid solution over ue gas ratio (L/G). Stripper pressure and eciency are never reported,

Fig. 5. Flow sheet of the ashed stripper conguration proposed by Barchas and Davis [1].

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

Table 2 Energy consumption of chemical absorption and stripping systems Authors Solvent Absorber Flue gas pressure loss composition (kPa) 13.8 14 6 34 20 39.2 20 Heat duty per ton recovered CO2 GJ=tCO2 Electric energy required per ton recovered CO2 kWh=tonCO2 13 20 80 48 120 73.5 73.9 72.3 62.9 63.1

Wiggins and Bixler [15] Dow FT-IL Pauley [12] Dow FT-IL Steinberg et al. [4] Dow FS-1 30% MEA Sander and Mariz [2] Econamine FG Suda and Iijima [5] Econamine FG This paper Yagi et al. [3] This paper 30% MEA 30% MEA

CO2=10% vol. O2=3% vol. 3.69 CO2=8.5% vol. O2=3.5% vol. 5.3 6 4.2 CO2 815% by mass 5.064 5.697 CO2=8% vol. O2=2% vol. CO2=8.55% by mole 4.066 Base plant 3.775 Pilot plant CO2=8.55% by mole 4.22 L/G=2.00 3.88 L/G=2.286 CO2=10% vol. 5.59 CO2=10% vol. 5.87 L/G=1.997 4.34 L/G=2.32

1907

1908

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

even though some indications can be helpful to understand their values. The comparison was then concentrated on global performance parameters, such as heat duty and power consumed per ton of recovered CO2. Table 2 shows a synopsis of the comparison with published data. It can be seen how overall performance parameters of the results obtained with the present model are consistent and in agreement with data published in the literature. A modied conguration of the capture system proposed by Barchas and Davis [1] was also studied (Fig. 5). A ash separator was located between the heat exchanger (HEAT-EX) and the stripper, with the aim of reducing its heat duty by extracting part of the CO2 from the rich solution before the regeneration process. The amount of CO2 extracted in the ash separator is about 18% of the total recovered CO2, but the benets of a smaller heat duty in the stripper are overcome by the larger amount of sensible heat necessary to heat the rich solution. The modied plant was simulated with a ue gas ow rate of 272 t/h, a mole concentration of 8.55% of CO2 (37 t/h), an L/G ratio of 2.0 and the same heat duty at the reboiler (39 MW), giving a slightly reduced covered CO2 ow rate. 6. CO2 liquefaction for transportation and storage Optimal conditions for liquid CO2 transportation dier depending on the means. For transportation in pipelines, the more appropriate parameters would be ambient temperature and 140 bar pressure. For transportation in ships, the optimal pressure and temperature are 6 bar and 508C, respectively, in order to apply the same technology as that used in LNG tanks [16]. After the stripper, the H2O and CO2 mixture is cooled to 258C to separate water from the CO2 before the liquefaction plant. This is necessary to avoid ice and clathrates formation. To liquefy CO2 at 140 bar and 258C, a four stage intercooled compression was used, with a pressure ratio in each stage of 2.97 (Fig. 6 and Table 3). To liquefy CO2 for ship transportation, the Hampson cycle was used, with four intercooled compression stages, which raise the pressure to 140 bar with subsequent throttling to 6 bar. The gaseous fraction after throttling is recirculated before the second stage at a temperature of 108C (Fig. 7 and Table 4).

Fig. 6. Liquefaction system for transportation in pipelines.

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915 Table 3 Results for the liquefaction plant for CO2 transportation in pipelines (P = 140 bar, T = 258C) Coal CO2 outlet ow rate (kg/h) H2O outlet ow rate (kg/h) Heat duty (MW) Power required (MW) Electric energy per ton liqueed CO2 kWh=tCO2 liq: 213,100 477 68.2 19.5 91.5

1909

Natural gas 130,000 291 40.9 11.9 91.5

7. Application of a CO2 removal and liquefaction system to a power plant The CO2 removal system described above was applied to a 320 MW conventional fossil fuel red steam power plant, with the following characteristics (Fig. 8 and Table 5): 1. 7 feedwater heaters, 4 on the high and medium pressure line and 3 on the low pressure line; 2. 1 deaerator located between the high and low pressure feedwater heaters; 3. Steam generator with high pressure at 150 bar and reheat at 35 bar. Two dierent fossil fuels were used in the simulations of the power plant: coal and natural gas. For each fuel, the ue gas composition for the CO2 removal plant was calculated, keeping the same assumptions valid in all the simulations and setting the optimal parameters previously found. The CO2 removal system was assumed to separate 90% of the CO2 contained in the ue gases (Table 6). The ue gas temperature at the cooling system inlet was assumed at 1108C. The ue gas is assumed to be available at the absorber inlet free of NOx and SOx, and that ue gas clean-up devices are present. The ue gases reach a temperature of 408C after being cooled in a direct contact heat exchanger consisting of a packed tower where the cooling uid is water at 258C. Cooling towers are modelled with two theoretical stages with Rushig rings packing and give a pressure loss of 0.1 bar on the gas side (Fig. 9). The cooled ue gases are then compressed at a pressure of 1.3 bar to balance the pressure losses in the absorber and the stripper. Cooling the ue gases from natural gas combustion to

Fig. 7. Liquefaction system for transportation in tank ships.

1910

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

Table 4 Results for the liquefaction plant for CO2 transportation in tank ships (P = 6 bar, T=508C) Coal CO2 outlet ow rate (kg/h) H2O outlet ow rate (kg/h) Heat duty (MW) Power required (MW) Electric energy per ton liqueed CO2 kWh=tCO2 liq: ) 213,100 1600 86.3 28.4 133 Natural gas 130,000 970 52.0 17.3 133

408C also reduces their ow rate because of the condensation of about 72 t/h of water. Instead, coal combustion ue gases are humidied in the cooling process (Table 7). Yagi et al. [3] calculated 300,000 m3/h as the maximum volume ow rate that could be treated in the absorber of a capture plant owing to economical limits of the size of the absorber. Three parallel trains consisting of an absorber, a heat exchanger and a stripper are, therefore, necessary to treat ue gases from natural gas combustion. For coal combustion, four trains are necessary due to the higher CO2 concentration (Table 8). The main changes to the steam power plant are as follows: 1. Reboilers are heated by means of steam extracted from the power plant. Steam is extracted at 5 bar from the same point as the deaerating steam, and the condensate is then sent back to the deaerator. This causes a signicant reduction of the steam ow rate in the condenser and in all the low pressure feed lines from the condenser to the deaerator (about 55 kg/s for the natural gas red plant and 93 kg/s for the coal red plant). Smaller steam ow rates are, therefore, necessary in the low pressure feedwater heaters. 2. The heat of condensation extracted at the top of the strippers can be used as a replacement

Fig. 8. Steam power plant ow sheet.

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915 Table 5 Base load performance of the power plant Net power output (MW) Eciency (%) Feedwater ow rate (kg/s) Cooling ow rate in the condenser (kg/s) Table 6 Characteristics of the ue gas Coal Gas ow rate at the stack (t/h) H2O (mole-fraction, %) CO2 (mole-fraction, %) N2 (mole-fraction, %) O2 (mole-fraction, %) Pressure at absorber inlet (bar) 1205 6.2 13.2 75.8 4.8 1.1

1911

320.9 44.3 284 19,290

Natural gas 1037 17.4 8.7 72.1 1.7 1.1

Fig. 9. Flue gas cooling and compression system. Table 7 Flue gas composition after compression (IN-ABS ow, Fig. 7) Coal Temperature (8C) Mass ow rate (t/h) Volumetric ow rate (m3/h) CO2 (mole fraction, %) H2O (mole fraction, %) Compressor power duty (MW) 66.5 1214 889,000 13.1 7.3 9.0 Natural gas 66.5 965 725,000 9.8 7.6 7.3

1912

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

Table 8 Main parameters of CO2 removal plant Coal Liquid to Gas ratio L/G CO2 inlet mass ow rate (t/h) Recovered CO2 mass ow rate (t/h) Overall inlet volume ow rate (m3/h) MEA make up (kg/h) Condenser temperature (8C) Condenser heat duty (MW) Reboiler heat duty (MW) Electric energy per ton recovered CO2a kWh=tCO2 rec: ) Heat duty per ton recovered CO2 GJ=tCO2 rec:
a

Natural gas 2.27 47.9 43.3 241,645 71.5 90.5 3.5 46.5 62.6 3.86

2.74 59.2 53.3 222,337 79.7 91.1 7 58.5 47.6 3.95

Including compressor and pump of ue gases cooling and compression system.

of the third low pressure feedwater heater, with 21 MW of heat transferred at 91.18C (Fig. 8).

8. Performance of a power plant with CO2 removal system After having calculated the performance of all the components of the CO2 removal and liquefaction system, the overall impact on the power plant can be determined (Table 9). Between the two alternatives of CO2 liquefaction, the optimal for pipeline transportation was considered because it has the lowest energy consumption per ton of recovered CO2. Removing CO2 from coal combustion gases has a bigger impact on power plant performance
Table 9 Impact of CO2 removal and liquefaction system on a steam power plant Base power plant Net power output (MW) Eciency (%) Cooling water ow rate DT=58C (kg/s) % CO2 recovered Water make up (t/h) MEA make up (kg/h) Steam extraction for the reboilers (kg/s) Cooling water per ton recovered CO2 kg=tCO2 rec: Water make up per ton recovered CO2 kg=tCO2 rec: MEA make up per ton recovered CO2 kg=tCO2 rec: % avoided CO2 320.9 44.3 19,285 Natural gas red power plant with CO2 recovery 267.2 36.9 23,338 90.4 214.5 55.5 112,270 1.65 88 Coal red power plant with CO2 recovery 237.3 32.7 22,115 90.0 108 318.8 93.1 47,800 508 1.50 86.5

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

1913

due to the higher CO2 concentration in the ue gases. Using natural gas has one more advantage, consisting of the elimination of water make-up to the CO2 removal plant, because water losses are completely covered by water recovery before the absorber and before the liquefaction system. MEA make-up has to balance the losses from evaporation at the absorber and stripper vents. Cooling water requirements for ue gas cooling increase the overall cooling ow rate by 20% in the case of natural gas combustion and by 16% in the case of coal combustion.

9. Cost analysis Having determined the performance and the characteristics of the components that are present in the CO2 removal and liquefaction plant, the capital and variable costs of the plant and their inuence on the cost of electricity were calculated. Dierent costs resulted for coal and natural gas red plants, and a detailed summary is reported in Table 10. The addition capital cost per kW electrical power after the insertion of a CO2 capture plant is 1617 US$/kWe for natural gas red power plants and 2011 US$/kWe for coal red power plants. Such high capital cost, together with the signicant reduction in the performance of the power plant, has a big inuence on the cost of electricity. Yearly maintenance costs were assumed to be 3% of the overall capital costs, whereas the MEA and water make-up were
Table 10 Capital costs of the CO2 removal and liquefaction planta Natural gas Flue gas cooling and compression Packed towers (Units) Heat exchangers (Units) Centrifugal compressors (Units) Overall cost including installation (US$) CO2 removal system Absorbers (Units) Strippers (Units) Heat exchangers (COOLER) (Units) Heat exchangers (LEAN-RICH) (Units) Reboilers (Units) Condensers (Units) Overall cost including installation (US$) Liquefaction Reciprocating compressors (Units) Heat exchangers (Units) Centrifugal compressors Overall cost including installation (US$)
a

Coal 3 (h = 10 m, d = 8.15 m) 3 1 106,175,000 4 (h = 47 m, d = 8.45 m) 4 (h = 40 m, d = 5.24 m) 8 4 12 4 421,080,000 1 7 3 117,930,000

3 (h = 10 m, d = 8.15 m) 9 1 114,328,000 3 (h = 47 m, d = 8.45 m) 3 (h = 40 m, d = 4.75 m) 6 3 6 3 296,289,000 1 6 3 108,296,000

h = height, d = diameter.

1914

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

Table 11 Additional costs of the kWh with the CO2 removal plant Natural gas Maintenance (US$/year] MEA consumption (t/year) Cost of MEA (US$/year) Water consumption (t/year) Cost of water (US$/year) Fixed costs (US$/year) Annual equivalent (10%; 15 years) (US$/year) Additional cost of kWh (US$) 15,565,000 1820 1,820,000 126,065,000 29,625,000 68,210,000 115,253,000 0.050 Coal 19,356,000 2750 2,750,000 88,960,000 20,906,000 84,822,000 127,838,000 0.0623

accounted separately by assuming operation for 8600 h/year and a cost of MEA and water of 1000 $/t and 0.235 $/t, respectively [7]. All the cost gures are accounted in 1998, with an investment life of 15 years at an interest rate of 10%. The cost of a kWh generated by a steam power plant with a CO2 removal and liquefaction system will be given by the sum of the cost of the kWh of the base power plant and the additional costs reported in Table 11. It can be seen that the cost of electricity more than doubles with reference to an average price of 0.0686 US$/kWh.

10. Conclusions The application of CO2 removal systems to power plants may greatly reduce CO2 emissions. A performance goal of 90% reduction can be achieved in conventional power plants and with normal CO2 concentration in the ue gases. The model described in this paper allows the estimation of optimal absorber and stripper performance parameters for low energy consumption. The goal of 90% CO2 emissions reduction is, however, penalised by the very high capital costs of the removal plant which make this solution not feasible at the current price of electricity. Reductions of cost can only be expected from a high volume of CO2 removal plant sales, because all the components are state of the art for the chemical industry, and the largest inuence on costs must be attributed to the absorber and stripping columns, which have a large size and require considerable construction and installation work.

Acknowledgements Partial funding of CNR is gratefully acknowledged.

U. Desideri, A. Paolucci / Energy Conversion & Management 40 (1999) 18991915

1915

References
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] Barchas R, Davis R. Energy Conversion and Management 1992;33(58):33340. Sander MT, Mariz CL. Energy Conversion and Management 1992;33(58):3418. Yagi T, Shibuya H, Sasaki T. Energy Conversion and Management 1992;33(58):34955. Steinberg M, Cheng HC, Horn F A system study for the removal, recovery and disposal of CO2 from fossil fuel power plant in the US 1984 DOE Report CH/00016. Suda T, Iijima M. Energy Conversion and Management 1992;33(58):31724. Wilson MA, Wrubleski RM, Yarborough L. Energy Conversion and Management 1992;33(58):32531. Corti A, Lombardi L, Manfrida G ASME Paper 98-GT-395 1998. Mathieu Ph ASME Paper 98-GT-383 1998. Wall G, Yantovskii E, Lindquist L, Tryggstad J Energy 1995;20:823828. Desideri U, Corbelli R. Energy Conversion and Management 1998;39(9):85767. Langeland K, Wilhelmensen K. Energy Conversion and Management 1993;34(911):80714. Pauley CR Chemical Engineering Progress May 5962 1984. Danckwerts PV. Chemical Engineering Science 1979;34:4436. Hagewiesche DP, Ashour SS, Al-Ghawas HA, Sandall OC. Chemical Engineering Science 1995;50(7):10719. Wiggins WW, Bixler RL. Energy Progress 1983;3(3):1324. Golomb D. Energy Conversion and Management 1997;38:27986.

Das könnte Ihnen auch gefallen