You are on page 1of 11

Modeling and Analysis

Reduction of greenhouse gas emissions by integration of cement plants, power plants, and CO2 capture systems
Luis M. Romeo, David Catalina, Pilar Lisbona, Yolanda Lara and Ana Martnez, CIRCE, Universidad de Zaragoza, Spain Abstract: Cement plants and power plants are two of the most signicant sources of greenhouse gases emissions. Many CO2 reduction options have been proposed in literature for both sectors. They are mainly focused on CO2 capture in power plants, but, in the short-term, industrial processes are going to play an important role in achieving this objective. In particular, one of the disadvantages in cement plants is that CO2 has two sources: fuel combustion and lime calcination. For this reason, proposed solutions could partially reduce a limited quantity of emissions. The sector is forced to use CO2 capture systems for further reductions. Preliminary results about the implementation of post-combustion and oxyfuel combustion systems for CO2 capture show low energetic penalties and important emissions reduction. Nevertheless, a detailed analysis, not only of CO2 emissions, but of raw materials and its disposal, primary energy and waste energy, could give optimum results from an environmental, energetic, and economic perspective. The combination/integration by industrial symbiosis of a power plant, a cement plant and a CO2 capture system is proposed in this work. Calcium-looping is chosen as the most suitable CO2 capture option for this application. The re-use of waste CaO coming from CO2 capture in the cement plant, and the utilization of waste energy from a clinker cooling and capture system to produce additional power are the main advantages of this proposal. Process ow diagrams and heat and mass balances are calculated and presented in this work. Results show a low value of the CO2 avoided cost, 12.4 /t, that is smaller than in any other combination of power plant with capture system or cement plant plus capture system, making this proposal economically very attractive. Moreover, an important amount of CO2 emissions is avoided 94% due to the energetic efciency augmentation, the reduction of raw and decarbonizated materials, and the CO2 capture system. 2011 Society of Chemical Industry and John Wiley & Sons, Ltd Keywords: calcium looping; cement industry; CO2 avoided cost; CO2 capture; greenhouse gas emissions

Correspondence to: Luis M. Romeo, CIRCE (Centro de Investigacin de Recursos y Consumos Energticos), Universidad de Zaragoza, Centro Politcnico Superior, Mariano Esquillor 15, 50018 Zaragoza, Spain. E-mail: luismi@unizar.es Received September 27, 2010; revised December 16, 2010; accepted December 17, 2010 Published online at Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/ghg3.005

72

2011 Society of Chemical Industry and John Wiley & Sons, Ltd | Greenhouse Gas Sci Technol. 1:7282 (2011); DOI: 10.1002/ghg3

Modeling and Analysis: CO2 capture in industrial sectors

LM Romeo et al.

Nomenclature
AC ASU CF CO2_capture CO2_ref FC FCF GCC LC TCR Avoided CO2 costs (/tCO2) Air Separation Unit Capacity factor () CO2 emissions in proposed case (Mt/year) CO2 emissions in reference case (Mt/year) Fuels cost (M/year) Fixed charge factor (year1) Grand Composite Curve Limestone cost (M/year) Total Capital Requirements (M)

Introduction
After electricity generation, the cement industry is a significant contributor to global CO2 emissions. According to IPCC1 and IEA2 between 5 and 7% of worldwide emissions are caused by the cement industry. Its considerable energy consumption, 45 GJ per tonne of cement,2 is mainly due to fuel combustion. Moreover, there is an additional CO2 released in the calcination reaction. The cement industry represents around 20% of industrial carbon emissions,3 with growing production that has gained importance in the last few years.3,4 For this reason, in the short term it will be an appropriate sector to apply important measures to reduce CO2 emissions. CO2 emission is in the range of 0.720.98 tonne CO2 per tonne of cement2 from which 5070% comes from calcination,58 4030% from fuel combustion, and around 10% from transportation and other ancillaries.5,8 These numbers could hardly be improved. The application of best available technologies may reduce the energy consumption to 3 GJ/t of cement.9 There are three main options to reduce CO2 emissions: energy efficiency, fuel switch, and the production of blended cements. The possibilities of implementing energy efficiency improvements to reduce energy input are scarce5,8,9 because of the the high efficiency of modern cement plants. Moreover, the improvements will represent a small reduction as fuel combustion is only responsible for 3040% of CO2 emissions. The same conclusion could be applied to the fuel switch. Although waste-derived alternative fuel has been extensively used in the cement industry, it might have adverse effects on the quality of the cement;5 it limits its application as well as the associ-

ated emissions reduction. The reduction of CO2 by the production of blended cement, using coal fly ash or blast furnace slag is also limited;5,8 further research on its effect on cement characteristics is required.10 Furthermore, the quality of these industrial byproducts is a serious handicap since carbon content in fly ash could make it inappropriate as a cement extender.11 Nevertheless, they could be suitable as CO2 adsorbers and partially capture cement or power plant emissions. Other possibilities to reduce the CO2 emissions associated with the cement industry includes the concrete carbonation (reabsorption of CO2 by cementitious materials due to carbonation)12 that requires further studies to quantify and analyze the variables that influence absorption; and mineral carbonation using the cement kiln dust (each tonne of clinker produces 0.150.20 tonnes of cement kiln dust)13 to fi x the CO2 previously released. None of these possibilities lead to serious reductions, although all of them are essential to deal with the challenge of greenhouse gas (GHG) reduction in the cement industry. For these reasons, several pieces of research have proposed CO2 capture as an attractive technology to be used in the cement industry. The combination of several of the aforementioned proposals to reduce emissions plus the separation of CO2 from flue gases seems to be the most attractive solution to solve the emissions associated with this industry. Post-combustion and oxyfuel combustion technologies have been proposed as possibilities for CO2 separation. Post-combustion based on amine scrubbing is suitable5,6,14 because CO2 concentration in the flue gases is higher than in power plants, therefore the separation efficiency could be larger and the capture costs could be reduced. Oxyfuel combustion has also been proposed5,6,15,16 because of the low oxygen consumption (compared with power plants) and the high concentration of CO2 due to calcination. A cost estimation has revealed that oxyfuel combustion is more economical than amine scrubbing 40/ tCO2 avoided for a 1 MT/y in a European cement plant6 which is an attractive cost in the medium term. Nevertheless, the most interesting option for cement plants is the option based on calcium-looping (carbonation-calcination of limestone) for CO2 capture.17,18 This concept has been proposed to power plants1921 showing low avoided costs.22,23 It has been also suggested for use in the cement industry under different schemes5,24,25 using the calcium-looping only

2011 Society of Chemical Industry and John Wiley & Sons, Ltd | Greenhouse Gas Sci Technol. 1:7282 (2011); DOI: 10.1002/ghg3

73

LM Romeo et al.

Modeling and Analysis: CO2 capture in industrial sectors

with the cement plant24 or integrating part of the waste heat in a Rankine cycle to produce power.25 As an advantage, the purge of deactivated CaO coming from the CO2 capture system could be used in cement plants as a raw material without additional CO2 release due to decarbonation of CaCO3. Although this purge is qualitatively low,26 quantitatively it is enough to satisfy the CaO requirements of a cement plant. This benefit, together with a proper integration of waste energy,27 could lead to important CO2 reductions, not only by capture emissions, but also by energy efficiency improvements and material re-use. All these strategies together could be defined as industrial symbiosis.28 Thus, the aim of the present paper is to propose and evaluate a novel scheme of industrial symbiosis between a power plant, a cement plant, and a CO2 capture system to re-use waste material and energy and overcome the problem associated with GHG emissions in two important industrial sources of CO2.

Description of the processes


Figure 1 shows the up-to-date processes used to produce cement and electricity. Cement plants need
Raw materials Fuel

mainly limestone and clay for clinker production. CO2 emissions are associated with fuel combustion and decarbonization of CaCO3. Power plants produce electricity with important CO2 emissions. Currently, the only interaction between both industries is the possibility of using fly ashes from power plants in cement production. But, as previously commented, the carbon content could negatively affect cement characteristics. The proposed system (Fig. 1) integrates both processes when a CO2 capture system is included to separate CO2 from the two installations. A CO2 capture system requires electricity from the power plant (for oxygen production and CO2 compression) and additional fuel for heating up and calcination of CaCO3. Outputs from this system are waste energy that could be integrated into the power plant or the cement plant increasing overall efficiency (reducing fuel consumption and CO2 emissions); and the purge that mainly consists of CaO could be used as input for the cement plant. It reduces limestone consumption and, as a consequence, reduces fuel input for the calcination of the saved limestone. Finally, the reduction of fuel consumption allows the integration of some waste heat from the cement plant in the power

Cement plant
Fly ashes

Clinker CO2

Fuel

Power plant

CO2 Electricity

Waste energy CaO (purge) Clinker Flue gas

Cement plant
CO2 Fly ashes CO2

CO2 capture

Raw materials Fuel Thermal energy Fuel

Fuel CaCO3

Power plant
Electricity Waste energy

CO2 to storage

Figure 1. Scheme of actual technology and proposal scheme for CO2 reduction with industrial symbiosis.

74

2011 Society of Chemical Industry and John Wiley & Sons, Ltd | Greenhouse Gas Sci Technol. 1:7282 (2011); DOI: 10.1002/ghg3

Modeling and Analysis: CO2 capture in industrial sectors

LM Romeo et al.

plant. Therefore, the proposed industrial symbiosis between a power plant, the cement plant, and a CO2 capture system of the two installations includes: Flue gases from the power plant and the cement plant are fed to the capture system to avoid CO2 emissions. Flue gases from the capture system will be mainly composed of nitrogen, moisture, and low amounts of oxygen and CO2. The purge from the CO2 capture process, mainly CaO, is used in the cement plant. It reduces the CO2 coming from the calcination of raw materials that accounts for 60% of the total emissions from a cement plant. This avoided CO2 is, in this scheme, captured in the oxyfuel calciner together with the CO2 released by coal. The surplus energy from the capture system is used to generate steam and to produce power in a turbine. Energy utilization should be thoroughly designed since efficiency maximization is not evident and forces the design of a new steam cycle to take advantage of the waste energy. Part of the waste energy released is used as coal, air, and oxygen preheating in the power plant, the cement plant, and the air separation unit (ASU). Therefore, the result will be a reduction in CO2 (avoided) due to overall efficiency improvements (preheating of streams, synergies between systems), a reduction of calcination emissions (by re-using the CaO), an augmentation of electricity production (reduction of specific CO2 emissions) and CO2 capture.

Power plant
Each system has been simulated independently to obtain the reference scenario. For comparison purposes and simplicity in future applications, the power plant is separated into two facilities with the same efficiency. The largest one will remain unchanged under every studied case, and the smallest one will integrate the waste heat from the capture system. In practice there would be only one power plant but, as we will describe, the CaO needed for a 3000 t/day cement plant corresponds to the CaO from the CO2 capture system of a 500 MWe power plant. The reference power plant consists of a boiler that produces live and reheated steam 600/600C at 180/50 bar and a steam cycle that produces 500 MWe. The ultimate analysis of the coal gives a composition of carbon 61.6%, oxygen 15.5%, hydrogen 4.9%, nitrogen 1.2%, moisture 10.1%, ash 6.7%, a low heating value 26.7 MJ/kg, and negligible sulfur content. The overall gross efficiency is 44.0%, thus 44.49 kg/s of coal are fed into the boiler. CO2 emissions sum up to 100.5 kg/s (specific emissions of 723.6 kg CO2/gross MWhe). The second power plant presents the same main variables, but produces 225.9 MWe. It is evident that the CO2 avoided cost depends on the reference plant; in this case the same efficiency has been assumed in both power plants in order to avoid any negative influence in the reference values.

Cement plant
The reference cement plant is illustrated in Fig. 2. It uses a dry process that requires 50% less energy than

Flue gases, 72.1 kg/s CO2, 29.8 kg/s

Drying & Grinding


Limestone, 3600 t/day Clays, 980 t/day 875C 875?C

Preheating
1050C 1050?C 1100C 1100?C

Air, 45.4 kg/s

Precalciner
Fuel, 81.8 MWth

900C 900?C Fuel, 53.8 MWth

Kiln

1450C 1450?C

Cooler
Clinker, 3000 t/day

Figure 2. Cement plant scheme.

2011 Society of Chemical Industry and John Wiley & Sons, Ltd | Greenhouse Gas Sci Technol. 1:7282 (2011); DOI: 10.1002/ghg3

75

LM Romeo et al.

Modeling and Analysis: CO2 capture in industrial sectors

the wet process, and consists of a raw material drier and grinder, a preheater, a precalciner, a kiln, and a final cooler before cement grinding. It is essential to detail the cement process simulation since it is necessary to assess the influence of the mass flow integration (CaO from capture system) on the cement plant streams and temperatures. For this reason mass and energy balances are performed in the main equipment of the cement plant. The cement plant production is assumed to be 3000 tonnes of clinker/day. According to Engin and Ari,29 energy losses from the main equipment of the cement plant account for 23.2%. This quantity was distributed among the preheater, calciner, kiln, and drying process. Kiln heat losses account for between 140 and 230 kJ/kg clinker;30,31 that represents between 5% and 8% of energy input (7% was assumed in this work). Preheater heat losses were assumed to be 5%, slightly higher than in Mujumdar et al.30 In the calciner and the drying processes, 6% and 5% respectively of energy loss were included to obtain an adequate amount of overall energy losses. The kiln is the key piece of equipment in a cement plant. The combustion excess air coming from the cooler at 1100C is 10%. Partially calcined limestone is fed at 900C. Total calcination and clinker formation are supposed to take place in this equipment. Energy balance allows calculating fuel and air input, flue gases mass flow and CO2 emissions (from coal and partial calcination). A clinker cooler is used to preheat combustion air that feeds both the kiln and the
Limestone, 3600 t/day Clays, 980 t/day Fuel, 135.6 MWth

precalciner. The temperatures of these streams are assumed to be 900C for precalciner and 1100C for the kiln. The output is the final clinker temperature, 133C a typical value for these installations. Calcination reaction was assumed to have an efficiency of 93% in the precalciner. Air for coal combustion is fed from the cooler heat exchanger at 900C, and flue gases from the kiln are introduced at 1100C. Flue gases reduce their temperature down to 1050C and are directed to the cyclonic preheaters, where the raw material is introduced at 325C and is led to the precalciner at 875C. Flue gases, at 634C, are led to the drying (raw material moisture is assumed to be 7%) and grinding process. Final flue gases temperature is around 290C. Results are in agreement with literature.2,6,7,30 To produce 1 t of clinker, 1.53 t of dry raw material (1.20 t limestone and 0.33 t clay) are needed. The energy input is 3.9 GJ/t clinker, 60.3% for precalciner consumption. It means that 61.5% of CO2 emissions are generated during the calcination and the total CO2 emissions are 0.86 t CO2/t clinker. Flue gases related to cement plant are 72.1 kg/s.

Overall system
The main energetic variables of the overall system are illustrated in Fig. 3. For comparison purposes the reference case comprises the 500 MWe power plant, the cement plant with a daily production of 3000 t, and an additional power plant of the same size as the

Cement plant

Clinker, 3000 t/day CO2, 29.8 kg/s

Fuel, 1188.2 MW th

Power plant

Electricity, 500 MW e CO2, 100.5 kg/s

Fuel, 513.4 MWth

Power plant

Electricity, 225.9 MW e CO2, 45.9 kg/s

Limestone, 3600 t/day Clays, 980 t/day Fuel, 1837.2 MWth

Clinker, 3000 t/day Electricity, 725.9 MW e CO2, 176.2 kg/s

Figure 3. Reference case. Energetic and materials streams, CO2 emissions.

76

2011 Society of Chemical Industry and John Wiley & Sons, Ltd | Greenhouse Gas Sci Technol. 1:7282 (2011); DOI: 10.1002/ghg3

Modeling and Analysis: CO2 capture in industrial sectors

LM Romeo et al.

Flue gases Q2

CO2 to storage Q3

Q1

CaCO3

CaO

CaO to cement plant CaCO3 Fuel

Flue gases from power plant and cement plant

ASU
Figure 4. CO2 capture cycle.

one designed to take advantage of the energy surplus from the CO2 capture system. As detailed below, the latter power plant has an electricity production of 225.9 MWe. Total fuel input is 1837.2 MWth. It is worthy to note that although cement plant fuel input represents only 7.6% of the total fuel requirements, CO2 emissions are 16.9% of the global system. Overall emissions are 176.2 kg/s (5.0 Mt/year) and electricity production is 725.9 MWe. These values correspond to the reference case for future comparisons.

CO2 capture plant


Figure 4 illustrates the CO2 capture process using a carbonation-calcination loop. It makes use of two interconnected circulating fluidized beds (CFB) to ensure a proper solid circulation and to enhance residence time and gas-solid contact. The flue gases from the existing power plant are sent to the carbonation CFB reactor. Due to sorbent deactivation, the extent of the carbonation reaction depends on the pretreatment of the solid, sorption properties of the raw material, percentage of purged solid, and CaO/ CO2 molar ratio.33 In capture systems for power plants, the minimization of tCO2 avoided cost is achieved with high CaO/CO2 molar ratio in the carbonation reaction and low purge percentages.26 Purge minimization limits the heat demanded in the calciner, the oxygen production cost, and the ancil-

larys consumption. A purge of 3.2% of the total solid inventory in the system is assumed. It is slightly larger than the optimum values obtained for CO2 capture in power plants26 but, in this case, the effect is counterbalanced since this stream will be re-used in the cement plant. The CaO/CO2 molar ratio is fixed in 4. The temperature of carbonator CFB operation must be set around 650C. The CaCO3-CaO solid stream leaving the carbonator is introduced into the calciner that operates at 950C to regenerate the sorbent. A significant energy input provided by coal oxyfuel combustion is required for calcination in the CFB reactor. An ASU is needed to provide oxygen, demanding considerable auxiliary consumption. A typical value of 220 kWh/ton O2 is assumed for ASU power consumption.1 A 2% of oxygen content in flue gases and a CO2 capture efficiency of 90% are assumed. Concentrated CO2 stream leaves the calciner at high temperature and its energy content may be used before the CO2 compression and conditioning stage. There is an important amount of waste energy released in the carbonation exothermic reaction (Q1) and contained in the clean flue gases (Q2), in the CO2 stream at very high temperature (Q3) and in the purge stream leaving the calciner at high temperature that could be used in a new steam power plant.22,26,27 Considering these data and the addition of the flue gases from both the power plant and the cement plant (100.5 and 29.8 kg/s of CO2), a coal mass flow of 33.15 kg/s and 63.55 kg/s of oxygen are required in the calciner. CaCO3 make-up stream is 29.4 kg/s and the purge is 16.5 kg/s of CaO. Additional CO2 from oxyfuel calciner is 87.8 kg/s, 74.9 kg/s associated with coal combustion and 12.9 kg/s from make-up calcination. Considering final temperatures of 150C, 336.9 MWth are released in the carbonator, 13.2 MWth is the waste energy from the solid purge and 251.2 and 225.4 MWth are associated with clean flue gases and CO2 stream before compression. To take advantage of this energy is essential to obtain an attractive integrated process.

Carbonator

Calciner

Process integration
To make an easier comparison, the surplus of energy from the CO2 capture system is used to generate steam in a new power plant without coal utilization. The size of this facility depends on the amount of CO2 captured, the additional ancillaries for oxygen production and CO2 compression, and the energy liberated in the carbonation, flue gases, and CO2 at high temperatures.

2011 Society of Chemical Industry and John Wiley & Sons, Ltd | Greenhouse Gas Sci Technol. 1:7282 (2011); DOI: 10.1002/ghg3

77

LM Romeo et al.

Modeling and Analysis: CO2 capture in industrial sectors

T (C)

1600 Hot streams before integration Integrated scheme 1400

1200

1000

800

600

400

200

0 0 200 400 600 800 Q (MW)

Figure 5. Grand composite curves.

Figure 5 shows Grand Composite Curves (GCCs) resulting from the pinch analysis. The solid line presents the pinch point of the system at high temperatures, and the possibility of steam cycle integra-

tion due to the great quantity of energy (837.7 MWth) available at high temperatures. This amount is composed of 684.4 MWth coming from the capture system, 132.4 MWth corresponding to the intercooling of the CO2 compression train, and 18.5 MWth from the final cooler at the cement plant. There is a surplus of energy in the final cooler at the cement plant because the saved fuel requires no combustion air. The dashed line in Fig. 5 shows the GCC of the integrated process. Most of the energy (779.4 MWth) is used in steam production in the new power plant, producing 343.3 MWe; 42.2 MWe of which are for oxygen production, 75.2 MWe for CO2 compression, and 225.9 MWe of electricity. Also, 7.5 MWth are used for oxygen and coal preheating, and 44.2 MWth in air preheating for both the cement and the reference power plant. Temperature level makes the integration within the system of the remaining 6.2 MWth impossible. The proposed industrial symbiosis between the power plant, the cement plant, and a CO2 capture plant treating the flue gases of the two systems is illustrated in Fig. 6. Each improvement causes a diminution in coal consumption, a direct CO2 reduction, and an indirect CO2 reduction in the rest of the system. Thus the waste energy used in the power plant reduces its fuel consumption and, then, the associated CO2 emissions of the overall system. Lime re-using reduces the cement fuel consumption for
Emitted CO2, 11.0 kg/s

CaO, 1249 t/day Limestone, 1379 t/day Clays, 980 t/day Fuel, 82.0 MWth

Captured CO2, 173.1 kg/s

CO2 capture

Cement plant

Clinker, 3000 t/day CO2, 14.3 kg/s Waste heat, 44.6 MW th

Fuel, 741.8 MWth Oxygen, 53.3 kg/s Limestone, 2221 t/day

Fuel, 1136.0 MWth

Power plant
Waste heat, 18.5 MW th

Electricity, 500 MW e CO2, 96.1 kg/s

Waste heat, 760.9 MW th Electricity, 343.3 MW e

Compression, Electricity, 75.2 MWe 42.2 MWe + 75.2 MWe Electricity, 225.9 MWe

Power plant

CO2, 0 kg/s

Limestone, 3600 t/day Clays, 980 t/day Fuel, 1959.8 MWth

Clinker, 3000 t/day Electricity, 725.9 MW e CO2, 11.0 kg/s

Figure 6. Proposed scheme of industrial symbiosis. Energetic and materials streams, CO2 emissions.

78

2011 Society of Chemical Industry and John Wiley & Sons, Ltd | Greenhouse Gas Sci Technol. 1:7282 (2011); DOI: 10.1002/ghg3

Modeling and Analysis: CO2 capture in industrial sectors

LM Romeo et al.

Table 1. Main variables in reference case and proposed integration.


Power plants
Coal (MWth) Oxygen (kg/s) Fresh CaCO3 (kg/s) Recycled CaCO3 (kg/s) CO2 combustion (kg/s) CO2 calcination (kg/s) Captured CO2 (kg/s) CO2 total (kg/s) 1701.6

Cement plant
135.6 41.7

Reference case
1837.2 41.7 157.9 18.3 176.2

Power plants
1136.0

Cement plant
82.0 16.0 14.0 6.9 7.4 14.3

CO2 capture
741.8 53.3 25.7 62.8 10.9 173.1 99.4

Integrated system
1959.8 53.3 41.7 165.8 18.3 173.1 11.0

146.4

11.5 18.3 29.8

96.1

146.4

96.1

calcination, its associated emissions, and the limestone descarbonization CO2 emissions. Table 1 includes a results comparison with the sum of the systems in the reference case and in the integrated system. Coal consumption increases 6.7% a reasonable value after considering the CO2 capture system. It is partly due to the important reduction (40.5%) of fuel input in the cement plant, from 135.6 to 82 MWth caused by the reduction of raw limestone substituted by CaO (61.7%). There is an important reduction of the emissions associated with cement production, from 29.8 to 14.3 kg/s (52%). As 90% of this quantity is captured later, the avoided CO2 in this plant is 95.3%. A 500 MWe power plant also reduces the fuel input and CO2 emissions by around 4.4%. For the global system, avoided CO2 emissions are 165.2 kg/s, around 94%. This value is higher than in any CO2 capture system and it is caused by the system symbiosis (fuel reduction, raw material, and energy re-use). Splitting the cause of each reduction: efficiency augmentation in power plant represents 4.4 kg/s of avoided CO2; material (CaO) re-use in cement plant represents 15.5 kg/s of CO2 saved, 11.0 kg/s due to calcination avoided and 4.5 kg/s due to fuel input for calcination; the rest of avoided emissions, 145.3 kg/s, is caused by the CO2 capture system. Evidently, captured emissions are larger than this quantity due to the additional energy requirements in the last system.

CO2 reduction is the cost of CO2 avoided.[1] This value provides a basis for comparison between a reference case and new proposals. The CO2 avoided cost (AC) is defined according to Eqns 1 and 2.

TCR * FCF
AC =

+ FC + LC CF CO 2 _ ref CO 2 _ capture

(1)

TCR = TCR carbonator + TCR calciner + TCR ASU (2) + TCR compression
To estimate the CO2 avoided cost for the proposed scheme of integration, the following costs are assumed: Specific total capital requirements (TCR) of power plants, 1100 /kW Specific TCR of cement plant, 160 /(t*yr) or 1900 /kWt24 Fixed charge factor (FCF), 0.1 Capacity factor (CF), 0.9 Fuels cost (FC), 1 /GJ Limestone cost (LC), 10 /t The same technology and specific costs are considered for both schemes. Calculations are done on the basis of extra total capital requirements and additional CO2 reductions. The largest additional cost is associated with the dual interconnected fluidized bed capital investment for the capture system. The re-use of purged CaO that reduces preheater, precalciner and kiln sizes implies a reduction of TCR of the cement plant. Since the estimation of the TCR reduction for a small system presents some difficulties, this cost reduction has not been considered in calculations as a conservative hypothesis.

CO2 avoided costs


Although both the reduction of raw material consumption and the improvement of energetic efficiency are highly influential issues for industrial symbiosis, the decision-variable that determines the feasibility of a proposed configuration is always economic. The most accepted variable to quantify any measures of

2011 Society of Chemical Industry and John Wiley & Sons, Ltd | Greenhouse Gas Sci Technol. 1:7282 (2011); DOI: 10.1002/ghg3

79

LM Romeo et al.

Modeling and Analysis: CO2 capture in industrial sectors

A 44.0% gross efficiency power plant, based on a thermal input of 779.4 MWth from the capture cycle, would produce 343.3 MWe and its cost rises up to 377.6 M. This cost is higher than the cost of the reference case power plant (225.9 MWe) since additional power for ASU and compression is required. Net power output is the same in both cases, 225.9 MWe, but power plant size is different. Considering that the boiler represents 40% of the TCR of a power plant, the cost of calciner would be 152.7 M. In the power plant which integrates the waste energy from capture plant, the heat exchangers substitute the combustion boiler. The cost of these heat exchangers (including carbonator) is assumed to be the same as the boiler they substitute. ASU and CO2 compression power requirements (42.2 and 75.2 MWe) make the gross power output of this power plant (343.3 MWe) larger than the reference case (225.9 MWe). Therefore, the TCR of this additional power plant is higher in the proposed scheme. ASU and compression costs are calculated as Rubin et al.34 with exchange rate $ to of 0.8. Therefore, specific cost is 291.4 /kWth, and TCR is 216.2 M. Compression TCR is 50.5 M. Main cost calculations are summarized in Table 2. Despite a TCR augmentation of 548.5 M, from 956.2 to 1504.7 M, and an annual operation cost increment of 3.4 M/year, the strong reduction of CO2 emissions compensates these additional cost. Thus, the CO2 avoided cost is 12.4 /t, lower than the cost obtained when applying the CO2 capture system to

power plants or cement plants. The synergies between these systems would allow a fuel reduction that leads to CO2 reductions additional to those reached in the capture system alone. The calciner cost is the variable with largest uncertainty. A sensitivity analysis reveals that an increment of 25% in its TCR only causes an increment in avoided cost of 0.8 /tCO2. The cost splitting in different systems and the huge emission reduction, mean that the final avoided cost is not strongly affected by any individual cost variation.

Conclusions
The proposed system integrates a cement plant and a power plant with a CO2 capture system to separate CO2 from the two installations if they are located in proximity. The surplus energy from the capture system is used to generate steam and to produce additional power. Part of this electricity is used to produce oxygen and for CO2 compression. Additional waste energy is integrated in the power plant increasing overall efficiency; the purge for the CO2 capture system is used as an input in the cement plant thus reducing the raw material consumption and the fuel input for the calcination of the saved limestone. It has been demonstrated that industrial symbiosis by a combination of two power plants, a cement plant and a CO2 capture system is a very interesting option to tackle the challenge of reducing emissions in two of the largest industrial sources of CO2. A possible drawback is that both installations have to operate together. The exchange of energy, solid material, and emissions among these systems propitiate a lower CO2 avoided cost than any other combination of power plant or cement plant with a CO2 capture system. Calculations have revealed an interesting low CO2 avoided cost, 12.4 /tCO2 , as well as 94% avoided emissions when these three systems are integrated. Despite these excellent results, further research is required to study the influence of deactivated CaO in clinker production and the effect of the sulfur and the CaSO4 formed in the capture system on the cement characteristics. Finally, it should be noticed that TCR increases significantly; a welldefi ned legal and economic framework is therefore necessary to ensure the success of this kind of symbiosis.

Table 2. Capital and operational costs. CO2 avoided cost calculations.


Reference plants
Cement plant Power plant (500 MWe) Power plant (225.9 MWe/ 343.3 MWe) Calciner ASU CO2 compression 157.7 M 550.0 M 248.5 M

Proposed symbiosis
157.7 M 550.0 M 377.6 M 152.7 M 216.2 M 50.5 M

TOTAL CAPITAL REQUIREMENT


Fuel cost Limestone cost

956.2 M
52.2 M/yr 11.8 M/yr

1504.7 M
55.6 M/yr 11.8 M/yr

TOTAL OPERATIONAL COST


CO2 emissions CO2 avoided cost

64.0 M/yr
176.2 kg/s

67.4 M/yr
11.0 kg/s 12.4 /tCO2

80

2011 Society of Chemical Industry and John Wiley & Sons, Ltd | Greenhouse Gas Sci Technol. 1:7282 (2011); DOI: 10.1002/ghg3

Modeling and Analysis: CO2 capture in industrial sectors

LM Romeo et al.

Acknowledgements
The work described in this paper was supported by the R+D Spanish National Program from Ministerio de Ciencia e Innovacin, MICINN (Spanish Ministry of Science) under project ENE2009-08246. Financial support for A. Martnez during her PhD studies was provided by the FPU programme of the Spanish Ministry of Science and Innovation.
References
1. Metz B, Davidson O, Bosch P, Dave R and Meyer L, Mitigation, in Intergovernmental Panel on Climate Change, Fourth Assessment Report, Climate Change 2007. Cambridge University Press, UK (2007). 2. IEA (International Energy Agency), Energy Technology Perspectives 2006: Scenarios & Strategies to 2050. IEA, Paris (2006). 3. Allwood JM, Cullen JM and Milford RL, Options for achieving a 50% cut in industrial carbon emissions by 2050. Environ Sci Technol 44:18881894 (2010). 4. The European Cement Association, Activity Report 2007. Available at: http://www.cembureau.be [accessed 13 January 2011] 5. Bosoaga A, Masek O and Oakey JE, CO2 Capture technologies for cement industry. Energy Procedia 1:133140 (2009). 6. Barker DJ, Turner SA, Napier-Moore PA, Clark M and Davison JE, CO2 capture in the cement industry. Energy Procedia 1:8794 (2009). 7. Gartner E. Industrially interesting approaches to low-CO2 cements. Cement Concrete Res 34:14891498 (2004). 8. Deja J, Uliasz-Bochenczyk A and Mokrzycki E. CO2 emissions from Polish cement industry. International Journal of Greenhouse Gas Control 4:583588 (2010). 9. The European Cement Association, Best Available Techniques for the Cement Industry: A contribution from the European Cement Industry to the exchange of information and preparation of IPPC BAT REFERENCE; Document for the cement industry (1999). 10. Roskovic R and Bjegovic D, Role of mineral additions in reducing CO2 emission. Cement Concrete Res 35:974 978 (2005). 11. Maroto-Valer MM, Lu Z, Zhang Y and Tang Z, Sorbents for CO2 capture from high carbon y ashes. Waste Manage 28:23202328 (2008). 12. Galan I, Andrade C, Mora P and Sanjuan MA, Sequestration of CO2 by concrete carbonation. Environ Sci Technol 44:31813186 (2010). 13. Huntzinger DN, Gierke JS, Kawatra SK, Eisele TC and Sutter LL, Carbon dioxide sequestration in cement kiln dust through mineral carbonation. Environ Sci Technol 43:19861992 (2009). 14. Ho MT, Allinson GW and Wiley DE, Comparison of MEA capture cost for low CO2 emissions sources in Australia. International Journal of Greenhouse Gas Control in press (2011). 15. Zeman F, Oxygen combustion in cement production. Energy Procedia 1:187 194 (2009). 16. Ghosal S and Zeman F, Carbon dioxide capture and storage technology in the cement and concrete industry. In Developments and innovation in carbon dioxide (CO2 ) capture and storage technology. Volume 1: Carbon dioxide (CO2 ) capture, transport and industrial applications. Ed. by Maroto-Valer MM. Woodhead Publishing Limited, Cambridge, UK (2010).

17. Blamey J, Anthony EJ, Wang J and Fennell PS, The calcium looping cycle for large-scale CO2 capture. Prog Energ Combust 36:260279 (2010). 18. Anthony EJ, Solid looping cycles: A new technology for coal conversion. Ind Eng Chem Res 47:17471754 (2008). 19. Abanades JC, Anthony EJ, Wang J and Oakey JE, Fluidized bed combustion systems integrating CO2 apture with CaO. Environ Sci Technol 39:28612866 (2005). 20. Abanades JC, Anthony EJ, Alvarez D, Lu D and Salvador C, Capture of CO2 from combustion gases in a uidized bed of CaO. AIChE J 50:16141622 (2004). 21. Romeo LM, Usn S, Valero A and Escosa JM, Exergy analysis as a tool for the integration of very complex energy systems: The case of carbonation/calcination CO2 systems in existing coal power plants. International Journal of Greenhouse Gas Control 4:647654 (2010). 22. Romeo LM, Abanades JC, Escosa JM, Pano J, Gimnez A, Sanchez-Biezma A et al., Oxyfuel carbonation/calcination cycle for low cost CO2 capture in existing power plants. Energ Convers Manage 49:28092814 (2008). 23. Abanades JC, Grasa G, Alonso M, Rodriguez N, Anthony E and Romeo LM, Cost structure of a postcombustion CO2 capture system using CaO. Environ Sci Technol 41:55235527 (2007). 24. Rodrguez N, Alonso M, Grasa G and Abanades JC, Process for capturing CO2 arising from the calcination of the CaCO3 used in cement manufacture. Environ Sci Technol 42:6980 6984 (2008). 25. Rodrguez N, Alonso M, Abanades JC, Grasa G and Murillo R, Analysis of a process to capture the CO2 resulting for pre-calcination of the limestone feed to a cement plant. Energy Procedia 1:141148 (2009). 26. Romeo LM, Lara Y, Lisbona P and Escosa JM, Optimizing make-up ow in a CO2 capture system using CaO. Chem Eng J 147:25225 (2009). 27. Lisbona P, Martnez A, Lara Y and Romeo LM, Integration of carbonate CO2 capture cycle and coal-red power plants. A comparative study for different sorbents. Energ Fuel 24:728 736 (2010). 28. Hashimotoa S, Fujita T, Gengc Y and Nagasawad E, Realizing CO2 emission reduction through industrial symbiosis: A cement production case study for Kawasaki. Resour Conserv Recy 54:704710 (2010). 29. Engin T and Ari V, Energy auditing and recovery for dry type cement rotary kiln systems A case study. Energ Convers Manage 46:551562 (2005). 30. Mujumdar KS, Ganesh KV, Kulkarni SB and Ranade VV, Rotary cement kiln simulator (RoCKS): Integrated modeling of pre-heater, calciner, kiln and clinker cooler. Chem Eng Sci 62:2590260 (2007). 31. Mujumdar KS and Ranade VV, Simulation of rotary cement kilns using a one-dimensional model. Chem Eng Res Des 84:165177 (2006). 32. IEA, Cement Technology Roadmap 2009: Carbon emissions reductions up to 2050. [Online]. IEA (2008). Available at: http://www.iea.org/papers/2009/Cement_Roadmap.pdf 33. Abanades JC, Anthony EJ, Wang J and Oakey JE, Fluidized bed combustion systems integrating CO2 capture with CaO. Environ Sci Technol 39:28612866 (2005). 34. Rubin ES, Yeh S, Antes M, Berkenpas M and Davison J, Use of experience curves to estimate the future cost of power plants with CO2 capture. International Journal of Greenhouse Gas Control 1:188197 (2007).

2011 Society of Chemical Industry and John Wiley & Sons, Ltd | Greenhouse Gas Sci Technol. 1:7282 (2011); DOI: 10.1002/ghg3

81

LM Romeo et al.

Modeling and Analysis: CO2 capture in industrial sectors

Dr Luis M. Romeo Dr Luis M. Romeo is Project Manager in CIRCE and a lecturer in Thermal System Simulation and Zero Emissions Technologies, Carbon Capture and Storage at the University of Zaragoza. His research is focused in energy integration and CO2 capture systems (oxyfuel and CFB looping). He holds a PhD in Mechanical Engineering.

Yolanda Lara Yolanda Lara is currently undertaking her PhD studies in Renewable Energies and Energetic Efciency. Her work in CIRCE is focused on CO2 capture cycles by means of CaO and its integration with power plants. She holds a degree in Mechanical Engineering from the University of Zaragoza.

Ana Martnez David Catalina Tomas David Catalina Tomas is studying for the European Renewable Energy Masters at CIRCE. He specializes in Energy, heat and uids technology and obtained his BSc in Mechanical Engineering at the University of Zaragoza in 2010. Ana Martnez is a research engineer at CIRCE and is currently reading her doctoral studies in Renewable Energies and Energetic Efciency. Her work covers calcium looping for CO2 capture and uidization in loop-seals. She holds a degree in Mechanical Engineering from the University of Zaragoza.

Pilar Lisbona Pilar Lisbona is a research engineer at CIRCE and teaching assistant in Thermal Engineering at the University of Zaragoza. She achieved her MSc in Chemical Engineering in 2002. Since 2003, she has also worked on fuel cell systems at Fraunhofer Institut UMSICHT (Germany) and the University of Perugia (Italy).

82

2011 Society of Chemical Industry and John Wiley & Sons, Ltd | Greenhouse Gas Sci Technol. 1:7282 (2011); DOI: 10.1002/ghg3

You might also like