You are on page 1of 11

Journal of Cleaner Production 32 (2012) 251e261

Contents lists available at SciVerse ScienceDirect

Journal of Cleaner Production


journal homepage: www.elsevier.com/locate/jclepro

Assessment of CO2 capture technologies in cement manufacturing process


Konstantinos Vatopoulos*, Evangelos Tzimas
European Commission, Joint Research Centre, Institute for Energy, P.O. Box 2, 1755 ZG Petten, The Netherlands

a r t i c l e i n f o
Article history: Received 29 July 2011 Received in revised form 5 March 2012 Accepted 7 March 2012 Available online 21 April 2012 Keywords: Cement Carbon capture and storage Oxy-combustion Post-combustion capture Calcium looping

a b s t r a c t
In this paper, an assessment of the viability of 3 CO2 capture technologies for the cement industry is performed; post-combustion absorptive capture (MEA) and oxy-combustion options are concepts already used by other industries and currently explored by the power sector; calcium looping postcombustion capture technology (CL) is an emerging technology that has not been assessed before in a comparative manner. The comparison is carried out in terms of specic energy consumption, CO2 footprint, CO2 capture energy penalty, raw material consumption and energy recovery potential. This has been achieved through the modelling of the integration of these process concepts with a reference cement plant. The results show that for the same capture efciency (85%), calcium looping has an advantage as the specic energy consumption increases by 18%. In the case of MEA the increase is 45%. CL also has considerably higher energy recovery potential, which can also further reduce its CO2 footprint. However, chemical looping demonstrates a higher complexity of integration with an existing cement plant. Oxy-combustion, though showing lower capture efciency (60%), results in lower specic energy consumption than the base case cement plant, which results to a negative CO2 capture penalty. These results contribute to the identication of the most suitable CO2 reducing strategy for the cement industry. 2012 Elsevier Ltd. All rights reserved.

1. Introduction Making the transition to a low-carbon economy is one of the key priorities of Europe today. European energy and climate policies have already been adopted to combat climate change through the reduction of greenhouse gas (GHG) emissions and to secure energy supply whilst maintaining the competitiveness of the European economy. These policies have set ambitious but necessary targets. By 2020, GHG emissions need to reduce by 20%, energy efciency to improve by 20% and the contribution of renewable energy to increase to 20% (European Commission, 2008); while the long term target of Europe is to reduce domestic GHG emissions by 80e95% by 2050 (European Commission, 2011). Meeting these targets will have major implications to economy and in particular the power sector and the energy intensive industries. Among the technological solutions with the potential to assist these sectors reduce their GHG emissions is the CO2 capture and storage (CCS). CCS has been identied as one of the prioritised technologies of the European Strategic Energy Technology Plan (SET-Plan), the European initiative to support the large scale

* Corresponding author. Tel.: 31 2 24 56 52 48; fax: 31 2 24 56 56 16. E-mail address: konstantinos.vatopoulos@ec.europa.eu (K. Vatopoulos). 0959-6526/$ e see front matter 2012 Elsevier Ltd. All rights reserved. doi:10.1016/j.jclepro.2012.03.013

deployment of low-carbon energy technologies through increased R&D and demonstration. The objective of SET-Plan for CCS includes further developing the technology to allow for its wide-spread use in all carbon intensive industrial sectors (European Commission, 2008). The cement industry is a signicant industrial GHG emitter. In 2009 the production of cement in the European Union was 200 million tonnes (Mt), about 7% of total world production. In Europe alone, 158 Mt of CO2 were emitted in 2008 from cement plants (JRC, 2010), which according to Eurostat corresponds to 38.5% of emissions from the industry or 3.2% of the total European CO2 emissions. A special characteristic of the cement industry, is the fact that in a modern cement plant, 60% of the CO2 emitted by a cement plant results from the calcination of limestone (Eq. (1)), 30% from combustion of fuels in the kiln and the remaining 10% from other downstream plant operations (Bosoaga et al., 2009; EU, 2008). Thus, measures like energy efciency improvements (use of energy efcient equipment, replacement of old installations, process modications, etc.), fuel switching to waste as alternative fuel, and, cement blending using industrial by-products (Worrell et al., 2001; Taylor et al., 2006; Bosoaga et al., 2009), can only help decrease the CO2 emissions associated with energy conversion. Signicant overall emission reductions can only be achieved with the application of CCS technologies. Process conditions appear

252

K. Vatopoulos, E. Tzimas / Journal of Cleaner Production 32 (2012) 251e261

to be favourable for the application of capture technologies, as the concentration of the CO2 from the calcination process is higher, in the range of 14%e33%, than in coal (12e14%) or gas red power plants (4%). CO2 capture technologies, currently considered for the power sector, such as post-combustion capture using amine scrubbing and oxy-combustion, have been proposed for CO2 capture in cement plants (EU, 2008; CSI, 2009; Hassan, 2005), while Hegerland et al. (2006) proposed post-combustion amine absorption as the best method for capturing CO2 in operating cement plants. Recent studies propose the calcium looping (CL) cycle as an option for CO2 capture in power plants (Shimizu et al., 1999; Abanades et al., 2004; MacKenzie et al., 2007; Bosoaga et al., 2009; Naranjo et al., 2011). Moreover, the Technology Task Force of the European Technology Platform for Zero Emission Power Plants has chosen the post-combustion carbonate cycle as one of the two highest priorities for future R&D as an emerging concept in carbon capture (ZEP, 2008). Integrating the CL capture unit of a power plant with a cement plant, by feeding the purged CaO stream into the cement kiln, has also been proposed. A study by Rodriguez et al. presented a dedicated process for capturing the CO2 emitted from the calcination process using CL (Rodriguez et al., 2008a, b). This process however results in an overall 57.2% capture efciency, since it does not take care of the CO2 emitted because of the fuel burning. Naranjo et al. (2011) conducted cement plant integration with CL capture unit that would treat all emissions form the plant. However, quantitative results of the proposed process have not been published. The objective of this article is to analyse and compare the technical performance of the 3 aforementioned CO2 capture options and to attempt identifying the most suitable CO2 capture technology that can be employed in the short and medium term to the existing or new cement plants. Furthermore, the key technological aspects of these technologies are well proven and have been used in a large scale in different industrial sectors. In particular, the paper is based on the modelling of the integration of these CO2 capture technologies with a reference cement clinker making process and estimates the corresponding specic energy consumption, CO2 emissions penalty and energy recovery potential. It is acknowledged however that the successful implementation of CO2 capture technologies in the cement industry will depend not only on the technical performance, but also on the economics and the legislative/market environment. These issues however are beyond the scope of this paper. On Section 2 an overview of the technologies considered in this paper is provided. Section 3 describes the modelling methodology followed in this analysis. Finally Sections 4 and 5 present the results and their analysis as well as the key conclusions of this paper. 2. Materials and methods 2.1. The clinker process In both the pre-calciner and the cement kiln, CaO is produced via the calcination of limestone (Eq. (1)). An additional product of this reaction is CO2.

gypsum and other additives to produce the cement mix. The ue gases of the kiln are directed back to the pre-calciner, in order to preheat the homogenised raw materials (limestone, clay etc.) and raise its temperature to the desired pre-calcination level (>900  C), and nally are emitted to the atmosphere. In the cement kiln, the desired high temperatures are achieved by the combustion of fossil or other unconventional fuels, and hence the CO2 emissions of the kiln depend of the fuel type used. The main fuel used in cement plants in Europe is pet-coke with 38.6% of market share in 2006, followed by coal (18.7%), waste fuels (17.9%) and pet-coke/coal mixes (15.9%). Other fuels used include fuel oil, lignite and natural gas (JRC IPCC, 2010). The most efcient kilns consume 2.5 MJ of fuel energy per kg of clinker (Kabir et al., 2010). In terms of electrical energy, approximately, 10 kWh/t of clinker is consumed by auxiliary equipment, e.g. conveyor belts, and in the cement bagging plant. Other energy demanding processes include processing of raw materials using a roller crusher, grinding in a vertical mill, a dry (short) pre-calciner kiln and nish grinding in a ball mill/separator. The energy consumption of each of these processes ranges from 66 to 107 kWh per tonne of cement (70e113 kWh/t clinker). The average overall electricity consumption in a cement plant is 145 kWh/tonne of cement (Worrell et al., 2001, 2000). The range of values highlights the variability of the cement manufacturing processes. 2.2. Oxy-combustion technology for the cement industry The oxygen boosted combustion in the cement kiln leads to reduced nitrogen content that does not have to be heated up. This energy can instead be used for the calcination process, which can lead to the reduction of ue gas volume or to an increase of kiln capacity. However, this is limited to 23e50% since there is a maximum acceptable concentration of oxygen (30e35% by volume) to avoid excessive damage to the cement kiln and to manage NOx emissions due to increasing thermal NOx formation (ECRA, 2009). Part of the ue gas is mixed with the O2 feed stream to raise the feed temperature and adjust the kiln temperature. The CO2 concentration in the kiln preheater and pre-calciner rises from 20 to 30% in a conventional process to over 80% when oxy-combustion is applied, with the rest of the gas consisting of air leak in N2, O2, and traces of NO2, SO2 and Ar. This has a major effect on the energy balance as well as the ratio between the energy content of the kiln ue gases and the energy needed for the chemical reactions of the kiln feed. Hence, the recirculation rate affects the combustion temperature. The largest bottleneck is the partial pressure of CO2 which affects the reaction type. At a low partial CO2 pressure, CaCO3 decomposition increases until reaction completion. Under high partial CO2 pressures, there is no reaction until a minimum threshold temperature is reached. If the temperature drops below that critical point, CaO promptly returns to CaCO3 through the carbonation reaction (inverse calcination) (Garcia-Labiano et al., 2002). To integrate the oxy-combustion technology into the clinker making process, an O2 supply unit and a CO2 purication facility (to enrich the CO2 stream and allow its transport and storage based on the current European legislation (EU, 2008)) are required. Both these plant components signicantly inuence the energy consumption. Research efforts focus on energy saving in O2 separation, with chemical looping technology currently under development and cryogenic air separation the most energy efcient technology for O2 production, with a power consumption of 200e240 kWh/tO2 (ECRA, 2009; IEA, 2008, 2010), although the recent developments that have reduced the energy requirement below 200 kWh/tonne O2, to around 160 kWh/tonne of O2 (CCJ, 2011).

CaCO3 s /CaOs CO2 g DH 178:2kJ=mol

(1)

The CaO together with silica (SiO2), iron oxide (Fe2O3) and alumina (Al2O3) form the cement clinker. The kiln operates at a temperature of 1400e1450  C which is required to complete the reaction of belite with CaO to form alite (Ca3O$SiO4), the characteristic constituent of Portland cement (JRC IPPC, 2010). The clinker is cooled down to around 120  C and is ground together with

K. Vatopoulos, E. Tzimas / Journal of Cleaner Production 32 (2012) 251e261

253

2.3. Post-combustion capture technology for the cement industry Post-combustion CO2 capture in the cement making process involves the separation of CO2 from the cement kiln ue gas. Several studies, trials and theoretical assessments about the potential application of CCS measures in the cement industry have been carried out (IEA GHG, 2008; ECRA, 2009; Nakanishi, 2009; Puyvelde, 2009; CSI, 2009; Sarlis and Shaw, 2008; Bosoaga et al., 2009; Naranjo et al., 2011). A qualitative comparison of post-combustion CO2 capture technologies for the cement industry by Naranjo et al. (2011) concluded that solvent-based capture is less suitable for a cement plant, as it is an energy intensive nature and requires the presence of hazardous materials on plant location. In addition, the cement industry has minimal experience with handling and processing liquid chemical processes operating liquid solvent-based systems (i.e. absorption columns). However, although solids-based technologies seem to offer less stringent process retrot and ue gas conditions compared to the other post-combustion CO2 capture technologies making it a good t for the cement industry, they would also require extensive retrot compared to a solvent-based CO2 capture technology for full CCS integration. 2.3.1. Amine scrubbing capture The main challenge for implementing this technology is the scale-up of the process. In contrast to the power sector, where many CCS pilot and demonstration projects are being launched (e.g. European Commission EEPR review 2010), the cement industry has not yet embarked into CCS pilot or demonstration projects. The results from the demo projects currently under way in the power generation sector are not directly transferable to the cement sector. A typical 600 MW power station yields a ue gas stream of 13% CO2. In terms of energy input, this translates to 0.052 kg CO2 per MJ of coal (or 1.5 kg of CO2 per kg of coal) (NETL, 2007). A typical cement plant with a capacity of 1 Mt/y would yield a ue gas stream of 30% CO2. In terms of energy input, this translates to 0.26 kg CO2 per MJ of coal input (or 6.8 kg CO2 per kg of coal input) (IEA GHG, 2008). In other words, a cement plant produces 5 times more CO2 in a 2.3 times more concentrated ue gas stream, when compared to a similar coal input power plant. Considering that approximately 3500e4500 kJ of heat are used to separate 1 kg of CO2 from the gas stream (ECRA, 2009), post-combustion amine absorption will have a much larger effect in the energy consumption of a cement plant, and may be proven to not to be as attractive option as in the power plant sector. Contaminants (such as SO2, NO2) in the ue gases of cement making process can be detrimental to the operation of the scrubber unit, as they would poison the absorption solvent. The concentration of SO2 in the inlet to the amine absorption process should be a maximum of 10 ppmv (at 6% O2) (IEA GHG, 2008), Similarly for NO2, it has been reported that the concentration in the ue gas should be limited to 20 ppmv (at 6% O2) (IEA GHG, 2008). More information on the post-combustion capture technology can be found in (Metz et al., 2005; CSLF, 2010; IEA, 2008). 2.3.2. Calcium looping process CL is a high temperature CO2 capture technology that makes use of the reversible calcination reaction (Eq. (1)) and consists of the calcium oxideecarbonate cycle using limestone to separate CO2 from cement kiln ue gas. A regenerable CaO sorbent e usually derived from limestone e is repeatedly cycled between the two vessels. In one vessel (the carbonator) carbonation of CaO occurs, and CO2 is captured in a circulating uidized bed carbonator operating between 600 and 700  C (Alonso et al., 2009, 2010; Blamey et al., 2010; Martinez et al., 2011). The solids stream (mainly CaCO3 formed and nonreacted CaO) goes to the calciner, where in an atmosphere of O2/

CO2 at temperatures over 900  C the CaCO3 decomposes into CaO which is recycled in the carbonator and a CO2 concentrated stream. In order to achieve a CO2 composition suitable for storage (>95% CO2) this calciner operates with pure oxygen. Therefore, CL installation always requires the integration with a pure O2 source. Although chemical looping for provision of O2 is a low energy technology that has been investigated extensively recently, an ASU is assumed to provide the oxygen in this study, as the latter is a proven technology that has been extensively installed in a commercial scale. High reactivity of the CaO used is kept by continuously replacing part of the non reactive CaO by fresh CaO. Thus, part of the calcinators exit solid stream containing the spent (un-reactive) sorbent is continuously purged. The purge stream constitutes mainly of CaO, SiO2 and CaCO3 and can be fed to the adjacent cement kiln. The SO2 produced in the kiln because of coke combustion is also captured, as it reacts with CaO to form CaSO4 (sulphation reaction). Recent experimental results (Ridha et al., 2012) showed that the presence of SO2 in the ue gas decreased the sorbent tendency for CO2 capture and lower temperatures would favour carbonation over sulphation. Experimental results on the integration of CL and cement production by Dean et al. (2011a, b) concluded that sulphation is acceptable if very large amounts of makeup limestone are used, for instance where the spent lime supplied to the cement kiln, where the levels of sulphate in the lime product must be minimised (Dean et al., 2011a). CL utilises technologies that have been demonstrated at large scale (Wiemr et al., 2007). Large (460 MWe) atmospheric and pressurised uidised bed systems have been in operation (Blamey et al., 2010). Dean et al. (2011a) have reported that several independent projects have been initiated in order to scale-up Calooping technology, including pilot plant trials with CO2 capture in the USA, Canada and Spain (up to 120 kWth); larger scale demonstrations are planned (2 MWth) and have demonstrated CO2 capture efciency levels of 80e90. Demonstration installations of 200 kWth at Stuttgart University, 1 MWth at Darmstadt University and 1.7 MWth at La Pereda power plant near Oviedo have started operation or are under construction (Dean et al., 2011b). Naranjo et al. (2011) concluded that there is a high synergy potential with a cement plant conguration in terms of: 1) use of spent sorbent for clinker and cement production onsite and/or offsite, 2) recovery of available waste heat for onsite power generation to offset additional power consumption, and 3) use of the same main plant fuel to operate the CO2 capture system. 2.3.3. CO2 compression and purication In oxy-combustion, the resulting CO2 stream is not of the necessary storage quality (usually between 80% and 90% CO2) a CO2 purication step is required in order to reach the necessary CO2 content (95%). Flue gas liquefaction or physical separation by compressioneliquefaction is the most economic solution for the further CO2 purication (Zeman and Lackner, 2006). The energy consumption is strongly affected by the air in leakage (i.e. percentage of non-condensable gases in the stream like N2, Ar, O2, NOx, SOx) where a 4.5% increase in energy consumption has been reported for every 1% increase in air leakage (ECRA, 2009). Other factors include the CO2/impurities molar ratio and the oxidiser purity. 3. Calculation The main assumptions of the modelling analysis of the aforementioned processes are presented in Table 1. The modelling of the upstream fuel preparation unit and downstream CO2 storage facility, grinding and nal cement mixing unit operations and CO2 compression unit are outside the scope of

254 Table 1 Main modelling assumptions. Capacity Pet-coke LHV Coke conversion Kiln temperature Minimum O2 content in kiln ue gases Clinker lime saturation factor (LSF) CO2 content in the stream for storage SO2 emissions limit NOx emissions limit Fuel used

K. Vatopoulos, E. Tzimas / Journal of Cleaner Production 32 (2012) 251e261

100 33,500 99.5 1450 3 0.98 >95 3.56 1.79 Petroleum coke

t/h clinker kJ/kg %  C % (v/v) % kg/tonne clinker kg/tonne clinker (5% moisture)

In terms of emission specications, the average EU 27 emission limits for SO2 and NOx for the production of cement reported in the 2010 IPPC BREF document on cement production (JRC IPPC, 2010). Detailed energy integration analysis of the studied processes is outside the scope of this paper. Therefore a pinch analysis was not conducted. However the integration targets for the CO2 capture processes were identied and the list of the available hot and cold stream is presented for each process (see Table 6). 3.1. Reference cement clinker making process For the reference clinker making process, the conguration proposed by the IEA GHG technical study was adopted (IEA GHG, 2008). A GIBBS-type reactor model, which assumes thermodynamic equilibrium is assumed for all the solidegas reactions, is used in Chemcad for the pre-calciner and the kiln. The cement kiln is assumed to operate in steady state conditions at atmospheric pressure and a temperature of 1450  C. All the solid/ gas separations and heat exchangers are modelled as ash vessels, to avoid unnecessary complexity in the model. 3.2. Oxy-combustion process The conguration of the oxy-combustion clinker making process was chosen based on the recommendations of the IEA GHG technical study (IEA GHG, 2008). This conguration captures the majority of the CO2 emissions from the cement plant by focussing only on the pre-calciner, where typically 60% of thee fuel is consumed and 95% of the calcination occurs. However, by using this conguration the complexity and uncertainties operating the kiln with pure oxygen as well as the inherent problem of air in leakage in the kiln are avoided. The process ow diagram consists of 7 process steps (see Fig. 1): A preheater where the O2 rich ue gas coming from the precalciner preheats the raw materials, a preheater where the kiln ue gases further preheat the raw materials, the cement kiln and clinker cooling units, and nally the FGD and CO2 purication units. A part of the CO2 rich ue gas stream is mixed with the O2 stream and returned back to the kiln in order to regulate the kiln temperature. The ratio between oxygen and recycled CO2 depends on the target kiln temperature. The heat integration target is to minimise fuel use in the cement kiln and pre-calciner. 6 integration points are identied:
-

Oxy-combustion O2 production energy Supplied O2 purity Clinker plant auxiliary electricity duty FGD CO2 purication and compression CL Precalciner temp Air-in leak Calcinator temperature Calciner thermal efciency Calcinator CaO/CO2 ratio (molar) CO2 purication and compression Amine scrubbing MEA/CO2 ratio in absorber (molar) O2 content in the ue gas before the absorber MEA quality Absorption column pressure Absorption feed temperature Stripper bottom temperature Absorber no. of trays Stripper no. of trays Stripper pressure CO2 purication and compression

576 95 522 7898 584

kJ/kg of O2 % (v/v) kJ/kg clinker kJ/SO2 kJ/kg CO2

>900 0.26 >950 93 5.05 538

 C kg air/kg clinker  C
%

kJ/kg CO2

1.79 >1.5 0.3 1.103 40 120 12 9 2.1 538

% v/v

bar  C  C

bar kJ/kg CO2

this modelling work, as they would add unnecessary complexity to the process ow modelling. However, the energy consumption of these units is taken into consideration in the analysis (see Table 1). The analysis focused on the most energy intensive part of the cement making process, which are the calcination, clinker formation and clinker cooling. Mass and energy balances have been calculated by process simulation using the commercial software Chemcad, according to the process congurations that were chosen for each case. The thermodynamic models (K-value and enthalpy models) used in simulation were chosen based on the chemical species present and process operating conditions (pressure, temperature etc.) and are based on thermodynamic equilibrium. For enthalpy models, Latent Heat enthalpy model and NRTL K-value model are used in main applications and Amine enthalpy and K-value models are used for the post-combustion CO2 capture unit operations. More information about these models and the selection process of thermodynamic package in Chemcad can be found in (Edwards, 2008). In order to guarantee that the produced clinker is of the same quality in all studied concepts, the Lime Saturation Factor was used. The LSF is a ratio of CaO to the other three main oxides. Applied to clinker, it is calculated based on Eq. (2):

The O2 feed stream is preheated with the oxy-precalciner; The O2 feed stream is further preheated by the kiln ue gas stream; The O2 feed stream is mixed with the hot recycle CO2 stream from the oxy-precalciner; The pet-coke feed stream is preheated by the oxy- pre-calciner ue gas; The preheater exit solid stream can be mixed and preheat the pet-coke feed to the oxy pre-calciner; The hot clinker can preheat the kiln air feed stream;

LSF

CaO 2:8*SiO2 1:2*Al2 O3 0:65*Fe2 O3

(2)

The LSF controls the ratio of alite to belite in the clinker. Typical LSF values in modern clinkers are 0.92e0.98 (WHD, 2011). In this article, an LSF 0.98 specication for the produced clinker was set.

Excess heat can be recovered as low pressure steam by further cooling the CO2 stream before the purication unit. The kiln ue gas can also be an input in a CHP plant, as there is a potential to cool down the stream prior to the FGD unit and produce both power and heat onsite. The modelling of the FGD is done with a GIBBS reactor model. The additional purication step is carried out by a FLASH separator model to remove water and a COMPONENT separator model, where the separation factors of each component is adjusted in order to match the ones reported in the literature (IEA GHG, 2008).

K. Vatopoulos, E. Tzimas / Journal of Cleaner Production 32 (2012) 251e261

255

Fig. 1. Block ow diagram of an integrated oxy-combustion clinker making process.

Fig. 2. Block ow diagram of an integrated amine scrubbing e clinker making process.

256

K. Vatopoulos, E. Tzimas / Journal of Cleaner Production 32 (2012) 251e261

3.3. Amine scrubbing process The process ow diagram presented in Fig. 2, was constructed based on the recommendations reported in the IEA GHG study (IEA GHG, 2008). Both the absorption and stripping columns are modelled using the SCDS rigorous multi-stage vapoureliquid equilibrium module, available in the database of the software. The thermophysical properties are modelled using the available Amine KenteEisenberg model, which is a simplied way to model the reactions (and phase equillibria) in a gas sweetening system. The variable that is set as an input in the model is the reboiler temperature at 120  C. The model calculates the column heat duty. Reboiler temperature cannot exceed 120  C to avoid amine degradation. Monoethanolamine (MEA) has been selected, based on its high absorption rate and as a proven concept with applications in the power generation industry (ECRA, 2009). The ue gas is treated to the acceptable NOx and SO2 levels, as mentioned in Paragraph 2.2.2, in an FGD unit prior to the amine absorption unit. The precalciner ue gas must be cooled down to 40  C before entering the absorption column, because the amine is very sensitive to temperature. Targeting in reducing pet-coke usage, 3 heat integration points have been identied:
-

amine scrubbing process, i.e. the regeneration of the amine in the desorber. 3.4. CL The chosen conguration (Fig. 3) serves the purpose of maximising the CO2 capture, by treating both the emissions from calcination and combustion. This process is similar to the one proposed by Naranjo et al. (2011), although the fuel here enters the process not only in the calcinator, but also in the cement kiln, in order to maintain the clinker formation temperature (1450  C). Furthermore, the limestone does not enter the process together with the other clinker raw materials, but as a separate feed stream in the calcinator, where it is calcined using pure oxygen in order to produce the CaO necessary both for the carbonator (CO2 capture) and the production of clinker. 7 process blocks were identied: raw materials preheater, pre-calciner, kiln and clinker cooler, carbonator calcinator and FGD. In the carbonator, the reaction of CaO with CO2 is limited by the 20% activity of the CaO in steady state (Abanades, 2002; Abanades and Alvarez, 2003; Rodriguez et al., 2010). Thus, a CaO/CO2 mole ratio of about 5 is assumed. All CO2 is released from CaCO3 and SO2 is completely captured by CaO. A GIBBS-type reactor is also used here to model the calcinator. The carbonator is modelled with a stoichiometric reactor in order to control the rates of the carbonation and calcination reactions and accommodate for the CaO reactivity limits (Hulin et al., 2000). The alternative of partly purging the carbonator product stream was investigated. The carbonator exit contains much higher number of CaCO3 and much less CaO (see Table 2). Thus, this option would increase the energy demand of the precalcinerekiln system for calcination. It would also result in recycling back in the kiln the CO2 that was captured in the carbonator. Therefore, this option was rejected.

The kiln pet-coke feed stream can preheated using the preheater hot gas exit stream; The pre-calciner pet-coke feed stream is preheated by mixing with the hot raw materials stream exiting the preheater; The MEA makeup stream is preheated by cooling the FGD exit gas stream.

The pre-calciner ue gas is cooled down by preheating the pet-coke stream remaining waste heat is used for steam production, which can partially cover the energy needs of the

Fig. 3. Block ow diagram of integrated CL e clinker making process.

K. Vatopoulos, E. Tzimas / Journal of Cleaner Production 32 (2012) 251e261 Table 2 Comparison of solid product streams of carbonator and calcinator. Temperature ( C) Mass composition % Ash CaCO3 CaO Pet-coke CaSO4 SiO2 Al2O3 Fe2O3 MgO 0.04 29.66 58.54 0.00 11.72 0.04 0.00 0.00 0.00 0.04 9.05 79.17 0.00 11.69 0.04 0.00 0.00 0.00 Carbonator exit 838.40 Calcinator exit 950.00 Table 4 Main CO2 capture performance indicators. Technology Base case - no Oxy-combustion capture 63% 929.7 14.7% 95.2% Amine scrubbing 85% 4952.2 7.3% 95.0% CL

257

Capture efciency CO2 capture energy penalty (kJ/kg CO2) CO2 in ue Gas 34.1% CO2 in CO2 stream e after cooling

85% 2806.6 11.4% 95.4%

Both the calcinator and the carbonator are modelled using GIBBS reactor Chemcad model. Heat integration target is the minimisation of pet-coke usage and 7 heat integration locations are identied:
-

The preheater ue gas stream is used to preheat the pet-coke, limestone and kiln air feed streams; The CO2 free ue gas exiting the carbonator preheats the limestone, pet-coke and oxygen streams; The concentrated CO2 stream exiting the calcinator is also used to preheat the solids feed of the calcinator.

Raw material consumption, in terms of kg of limestone per kg of clinker; -Energy consumption, by means of energy consumed per kg of clinker produced; -Energy recovery potential, by measuring the potential recovery as high pressure steam (200 bar) for power generation but also for low pressure steam for heat integration within the process; - CO2 footprint and the reduction in direct CO2 emissions by implementing each technology, by measuring the CO2 emitted per kg of clinker produced; - CO2 capture energy penalty when compared to the consumption of the base case process.
-

The remaining waste heat can produce high pressure (200 bar) steam for electricity production and low pressure (3.6 bar) steam. The high pressure steam is of particular interest as it can partly satisfy the energy needs for O2 production. 4. Results and discussion 4.1. Selection of performance indicators The comparison of the performance of the three capture technologies was made based on ve key indicators:

The mass balance modelling results for the studied processes (see Table 3) indicate that oxy-combustion process requires 2.2 times less input mass streams (2.4 times less cooling air stream and 32% less pet-coke stream). This difference was expected, as the oxy-combustion process has less diluted gas streams, due to the absence of N2. Furthermore, it captures less CO2 compared to the other 2 capture concepts. The main capture performance results (see Table 4) show an advantage of the CL process over amine scrubbing, in terms of energy penalty for CO2 capture. Oxycombustion treats less portion of the CO2 and consequently has less energy penalty (in this case negative since this process is less energy intensive than the base case) and higher concentration of CO2 in the ue gases.

Table 3 Mass balance modelling results. Technology Stream summary Input streams Raw materials mix CaCO3 Pet-coke Air Air in leak O2 Water MEA makeup Sum Base case e no capture kg/h 35,953 115,156 9037 4,015,541 26,400 0 455 0 4,202,543 36,912 117,676 9246 1,655,284 26,400 15,064 13,809 0 1,874,392 35,953 115,156 9037 4,015,540 26,400 0 455 168 4,202,709 37,160 111,999 14,732 4,015,540 26,400 19,740 136 0 4,225,708 Oxy-combustion Amine scrubbing CL Base case e no capture kg/kg clinker 0.35 1.12 0.09 39.10 0.26 0.00 0.00 0.00 40.92 0.36 1.15 0.09 16.12 0.26 0.15 0.13 0.00 18.25 0.35 1.12 0.09 39.11 0.26 0.00 0.00 0.00 40.93 0.36 1.09 0.14 39.25 0.26 0.19 0.00 0.00 41.31 Oxy-combustion Amine scrubbing CL

Output streams Clinker Surplus air Flue gas Water out Gypsum CO2 to storage Inert gases Sum CO2 emissions

102,694 3,823,449 276,352 0 47 0

102,693 1,522,850 180,744 1868 1803 47,362 17,072 1,874,390 26,506

102,685 3,821,047 192,089 51 86,838

102,295 3,908,454 120,278 1715 20 92,945

1.00 37.23 2.69 0.00 0.00 0.00 0.00 40.92 0.92

1.00 14.83 1.76 0.02 0.02 0.46 0.17 18.25 0.26

1.00 37.21 1.87 0.00 0.00 0.85 0.00 40.93 0.14

1.00 38.21 1.18 0.02 0.00 0.91 0.00 41.31 0.13

4,202,542 94,177

4,202,709 14,127

4,225,707 13,760

258 Table 5 Energy consumption results. Technology Energy consumption Fuel Electricity clinker plant ASU MEA scrubber FGD CO2 purication CO2 compression Sum Base case e no capture kJ/kg clinker 2918.08 522.00 0.00 0.00 0.00 0.00 0.00 3440.08 2985.67 522.00 84.49 0.00 8.25 212.19 43.90 3856.51

K. Vatopoulos, E. Tzimas / Journal of Cleaner Production 32 (2012) 251e261 Table 6 Heat duties and temperatures of streams targeted for heat integration. Oxy-combustion Amine scrubbing CL Clinker conventional Hot Hot clinker Hot Kiln ue gases Cold Raw materials Cold Kiln pet-coke feed Cold Excess air Cold Kiln air feed Tin ( C) Tout ( C) F (kg/h) 1450 1450 25 25 25 71 70 110 869 815 71 735 Q (MJ/h)

2918.33 522.00 0.00 3508.18 37.82 191.63 241.02 7418.98

4874.65 522.00 111.15 0.00 11.06 206.78 260.07 5985.70

102,694 358,546 276,352 481,405 151,109 533,738 9037 6068 4,015,541 188,594 136,854 92,779 18771

4.2. Raw material consumption The specic raw material consumption of the reference (no capture plant) is 1.5 kg per kg of clinker produced. The application of CO2 capture technologies does not affect raw material consumption. The small increase (3%) of specic consumption in the case of oxy-combustion can be attributed to the changed conditions in the cement kiln, where the reactions in the solid phase slow down, due to the increased concentration of CO2 in the gas phase. 4.3. Energy consumption As it can be seen form the specic energy consumption results (Table 5), by integrating the cement plant with a post-combustion CO2 capture process, the specic energy consumption increases by 45% for the case of MEA, 18% for the CL and decreases by 12% when oxy-combustion is applied (Fig. 4). In the MEA case the increased energy consumption comes from the heat demand of the desorber reboiler. The increase in energy consumption when a CL unit is integrated stems mainly from the additional fuel required to operate the calcinator and the additional O2 produced in the ASU to inject into the calcinator. Oxycombustion, including the ASU consumption, resulted in

Oxy-combustion Hot Hot clinker 1450 Hot Kiln ue gases 1450 Hot Oxy-precalciner ue gases 995 Cold Pet-coke kiln feed 25 Cold Pet-coke oxy-precalciner feed 25 Cold Kiln air feed 70 Cold Excess air 25 Cold Oxygen 25 Cold Raw materials 25

70 110 110 900 954 819 70 912 954

102,693 358,542 180,744 315,098 130,229 150,981 4940 3674 4306 3400 132,434 74,395 1,655,284 75,829 15,064 12,186 154,589 601,021 54117

CL Hot Hot Hot Cold Cold Cold Cold Cold Cold Cold

Hot clinker Kiln ue gases Calcinator ue gases Pet-coke kiln feed Pet-coke calcinator Excess air Kiln air feed Oxygen Raw materials CaCO3

1450 1212 950 25 25 25 70 25 25 25

70 130 110 1192 837 70 820 700 984 837

102,295 357,153 120,278 169,183 94,394 104,188 5737 5691 8995 6208 4,015,540 135,524 107,086 60236 19,740 14,470 37,160 149,210 111,999 74,664 184520

Amine scrubbing Hot Clinker Hot Kiln ue gas Cold Raw materials Cold Pet-coke kiln feed Cold Excess air Cold Kiln air feed Cold MEA makeup

1450 1450 25 70 25 70 25

70 110 869 815 70 820 46

102,685 358,514 278,927 489,628 151,109 533,738 9037 5723 4,015,540 140,945 194,493 149,079 168 13 18644

Fig. 4. Energy intensity in kJ/kg of clinker produced.

K. Vatopoulos, E. Tzimas / Journal of Cleaner Production 32 (2012) 251e261

259

Fig. 5. Energy recovery potential of the studied concepts.

a decrease in energy consumption, as the energy consumed in the ASU for the supply of O2 is less than the savings in fuel consumption. In conclusion, oxy-combustion clinker production is more less energy intensive than the base case process. CL CO2 emission reduction on yields the lowest energy penalty than amine scrubbing. 4.4. Energy recovery potential Table 6 presents the hot and cold streams targeted for heat integration for the studied concepts. As energy recovery potential for each of the studied concepts, the portion of the energy content of the exit streams, that after passing through the heat integration positions identied in Section 3 could potentially be used for power generation (200 bar steam) or heat production (3.6 bar steam) is considered (see Fig. 7). The heat requirements for the cold streams can be covered by the available heating from cooling the hot streams, and there is waste heat available in all cases, with CL demonstrating the highest potential. The CL process has 3.8 times higher energy recovery potential than the oxy-combustion and 11.5 times than amine scrubbing concept. This stems from the resulting high temperature CO2 and

ue gas streams. The power generation potential could cover the energy of the O2 production unit, and the auxiliary clinker production units electricity duty (Fig. 5). The results for the oxycombustion case also show a potential for power generation from waste heat, which could also cover the O2 production unit in this case. The available waste heat in the amine scrubbing process (125 kJ/ kg clinker) can be recovered as low pressure steam and satisfy 4% of the energy requirements of the amine regeneration boiler (3508 kJ/ kg clinker). This is quite lower than the 15% reported by ECRA (2009). 4.5. CO2 footprint This factor presents the CO2 that is not captured by one of the studied technologies. In this analysis only the direct avoided CO2 emissions are considered. The specic direct emissions for the modelled concepts, in kg of CO2 emitted per kg of clinker produced, do not include emissions from fuel use in supporting units (i.e. ASU, CO2 purication, steam boilers etc.). Emissions avoided by utilising the high grade waste heat for producing electricity should also be taken into account for a fair comparison between the different

Fig. 6. Specic CO2 emissions of the studied concepts.

Fig. 7. Specic CO2 capture energy penalty.

260 Table 7 Comparison with literature. Source Process concept

K. Vatopoulos, E. Tzimas / Journal of Cleaner Production 32 (2012) 251e261

Energy consumption kJ/kg clinker

CO2 capture energy penalty kJ/kg CO2 0 0 0 4705 4572 931 906 2831 1629

CO2 footprint kg CO2 emitted/kg clinker 0.92 0.85 0.82 0.14 0.17 0.26 0.31 0.15 0.38

CO2 capture rate kg captured/kg produced 0% 0% 0% 85% 85% 63% 62% 100% 58%

This paper IEA GHG, 2008 Rodriguez et al., 2008a, b This paper IEA GHG, 2008 This paper IEA GHG, 2008 This paper Rodriguez et al., 2008a, b

Base case e no capture Base case e no capture Base case e no capture Amine scrubbing Amine scrubbing Oxy-combustion Oxy-combustion CL CL

3440 3317 3158 7419 7256 3848 3780 5986 4000

technologies. However, since detailed energy integration is outside the scope of this paper, converting the energy recovery potential into avoided CO2 emissions stands also outside the boundary limits of this analysis. CL and amine scrubbing reduce the specic CO2 emissions by a factor of 6.8 and 6.7 respectively (Fig. 6). In the case of oxycombustion this factor is 3.6. This is a result of the lower capture efciency of the oxy-combustion conguration chosen. 4.6. CO2 capture energy penalty The CO2 capture energy penalty is the additional energy required as an input to the clinker process in order to capture the CO2. This is calculated using the following equation:

in the IEA GHG report stems form the fact that the energy consumption of the auxiliary units is much more detailed than in this study. 5. Conclusions This paper presents a preliminary comparative analysis of the 3 most CO2 promising technologies for implementation in the cement industry in the short and medium term. Oxy-combustion, post-combustion amine scrubbing and CL. These 3 technologies were analysed and assessed against 5 performance criteria; raw material consumption, specic energy consumption, CO2 capture penalty, CO2 footprint and specic energy recovery potential. CL technology demonstrates a conceptual advantage in the integration with a cement plant over the amine scrubbing process. While demonstrating a higher complexity of integration and smaller impact in reducing the CO2 footprint of the clinker making process, it has lower specic energy consumption and capture energy penalty than the amine scrubbing technology, at 85% capture efciency. In addition, CL technology shows also a much higher potential for recovering the waste heat produced, especially in the form of high pressure steam which can be used for power generation, further reducing its CO2 footprint. The complexity of the integration and the waste heat recovery potential need to be further analysed. Oxy-combustion technology reduces the overall energy consumption and results in a negative CO2 capture penalty of a cement plant. However, the conguration chosen here avoids the major kiln modications that would be required if the kiln was operating in an O2/CO2 atmosphere. This has the immediate effect of capturing only 60% of the produced CO2. The effects of kiln operation at an O2/CO2 environment must be completely understood in order to evaluate a conguration with increased capture efciency. Oxy-combustion and CL should be assessed further in a comparative manner, and for capture efciencies of 85% or more. Still, when comparing the oxy-combustion results of this study with the CL process proposed by Rodriquez et al. (2008a, b), which also has a 60% capture efciency (see Table 5), it can be concluded that if this capture efciency is accepted, oxy-combustion is still more favourable than CL. Acknowledgements This work has been carried out within the multi-annual work programme of the Assessment of Energy Technologies and Systems (ASSETS) Action of the European Commissions Joint Research Centre. Any interpretations or opinions contained in this

ECapture

EX ENC FCO2 ;caotured

(3)

where: Ex is the total energy consumption of process concept X, ENC is the total energy consumption of the base case and, FCO2 is the mass rate of the CO2 captured in process X. Amine scrubbing is the most energy intensive capture alternative, with 43% higher specic energy consumption compared to CL (see Fig. 7). Oxy-combustion is 81% less energy intensive, when compared to amine scrubbing. This stems not only from the lower capture efciency of the selected process in comparison to the CL and amine scrubbing concepts, but also to the fact that the oxycombustion clinker making process results in less fuel consumption per ton of clinker. 4.7. Literature comparison In order to validate the process ow models developed for each of the studied process concept, the specic energy consumption of each process (MJ/kg clinker) was compared to the specic energy consumption reported for similar process concepts reported in previous studies (IEA GHG, 2008; Rodriguez et al., 2008a, b). The results show that the models developed in this study result in energy consumption gures quite similar to the ones reported elsewhere (see Table 7). The higher energy consumption (4e5%) in the base case concept can be attributed to the lower level of energy integration in this study, compared to the literature studies. In the case of CL, the higher (50%) specic consumption stems form the different process conguration chosen, which aims at 86% CO2 capture rate, compared to the process presented by Rodriguez et al, where only 58% of the produced CO2 is captured. The higher (2%) consumption in the amine scrubbing and oxycombustion processes when compared to the numbers reported

K. Vatopoulos, E. Tzimas / Journal of Cleaner Production 32 (2012) 251e261

261

paper are those of the authors and do not necessarily represent the view of the European Commission. References
Abanades, J.C., 2002. The maximum capture efciency of CO2 using a carbonation/ calcination cycle of CaO/CaCO3. Chem. Eng. J. 90 (3), 303e306. Abanades, J.C., Alvarez, D., 2003. Conversion limits in the reaction of CO2 with lime. Energy Fuel 17 (2), 308e315. Abanades, J.C., Rubin, E.S., Anthony, E.J., 2004. Sorbent cost and performance in CO2 capture systems. Ind. Eng. Chem. Res. 43 (13), 3462e3466. Alonso, M., Rodrguez, N., Grasa, G., Abanades, J.C., 2009. Modelling of a uidized bed carbonator reactor to capture CO2 from a combustion ue gas. Chem. Eng. Sci. 64, 883e891. Alonso, M., Rodrguez, N., Gonzlez, B., Grasa, G., Murillo, R., Abanades, J.C., 2010. Carbon dioxide capture from combustion ue gases with a calcium oxide chemical loop. Experimental results and process development. Int. J. Greenh. Gas Control 4 (2), 167e173. Blamey, J., Anthony, E.J., Wang, J., Fennell, P.S., 2010. The calcium looping cycle for large-scale CO2 capture. Prog. Energy Comb. Sci. 36 (2), 260e279. Bosoaga, A., Masek, O., Oakey, J.E., 2009. CO2 capture technologies for cement industry. Energy Procedia 1, 133e140. Carbon Capture Journal, 2011. Air Separation Units for Coal Power Plants. Published on June 22 2011, available at: http://www.carboncapturejournal.com/ displaynews.php?NewsID805. CSI, 2009. WBCSD/ECRA Technology Paper on the Development of State of the Arttechniques in Cement Manufacturing: Trying to Look Ahead. CSLF, 2010. CSLF Technology Roadmap. Dean, C.C., Dugwell, D., Fennell, P.S., 2011a. Investigation into potential synergy between power generation, cement manufacture and CO2 abatement using the calcium looping cycle. Energy Environ. Sci. 6, 2050e2053. Dean, C.C., Blamey, J., Florin, N.H., Al-Jeboori, M.J., Fennell, P.S., 2011b. The calcium looping cycle for CO2 capture from power generation, cement manufacture and hydrogen production. Chem. Eng. Res. Des. 89, 836e855. Edwards, J.E., 2008. Process Modelling Selection of Thermodynamic Methods. http://www.chemstations.com/content/documents/Technical_Articles/thermo. pdf. The EU technology platform for zero emission fossil fuel power plants, recommendations for RTD, support actions and international collaboration activities within FP7 energy work programmes in support of deployment of CCS in Europe, http://www.zeroemissionplatform.eu/website/docs/ETP%20ZEP/TTech %20Input%20FP7%203rd%20call%20080418.pdf, 2008. European Cement Research Academy, 2009. ECRA CCS Project e Report about Phase II. Technical Report TR-ECRA-106/2009. European Commission, 2008. Communication from the Commission to the Council, the European Parliament, the European Economic and Social Committee of the Regions: 20 20 by 2020 e Europes Climate Change Opportunity. COM, p. 30. European Commission, 2011. Communication from the Commission to the Council, the European Parliament, the European Economic and Social Committee of the Regions: a Roadmap for Moving to a Competitive Low Carbon Economy in 2050. COM, p. 112. European Commission DG Joint Research Centre (JRC), 2010. Scientic and Technical Report on Energy Efciency and CO2 Emissions: Prospective Scenarios for the Cement Industry EUR 24592 EN. European Commission DG Joint Research Centre JRC IPPC, 2010. Reference Document on Best Available Techniques in the Cement, Lime and Magnesium Oxide Manufacturing Industries. Adopted May 2010. Garcia-Labiano, F., Abad, A., Diego, L.F., Gayan, P., Adanez, L., 2002. Calcination of calcium based sorbents at pressure in a broad range of CO2 concentrations. Chem. Eng. Sci., 2381e2393.

Hassan, S.M.N., 2005. Techno-economic Study of CO2 Capture Process for Cement Plants. University Waterloo, Ontario, Canada. Hegerland, G., Pande, J.O., Haugen, H.A., Eldrup, N., Tokheim, L., Hatlevik, L., 2006. Capture of CO2 from a Cement Plant Technical Possibilities and Economic Estimates, GHGT 8, Norway. Hulin, L., Guangbo, Z., Rushan, B., Yongjin, C., Gidaspow, C., 2000. A coal combustion model for circulating uidized bed boilers. Fuel 79, 165e172. International Energy Agency Greenhouse gas R&D Programme (IEA GHG), 2008. CO2 Capture in the Cement Industry. Report 2008/3. International Energy Agency Energy Technology Systems Analysis Programme, 2010. Technology Brief I03 e Cement Production. Kabir, G., Abubakar, A.I., El-Nafaty, U.A., 2010. Energy audit and conservation opportunities for pyroprocessing unit of a typical dry process cement plant. Energy 35, 1237e1243. MacKenzie, A., Granatstein, D.L., Anthony, E.J., Abanades, J.C., 2007. Economics of CO2 capture using the calcium cycle with a pressurized uidized bed combustor. Energy and Fuels 21 (2), 920e926. Martinez, I., Murillo, R., Grasa, G., Rodriguez, N., Abanades, J.C., 2011. Conceptual design of a three uidised beds combustion system capturing CO2 with CaO. Int. J. Greenh. Gas Control 5 (3), 498e504. Metz, B., Davidson, O., De Coninck, H., Loos, M., Meyer, L., 2005. Carbon Dioxide Capture and Storage. Summary for Policymakers and Technical Summary. IPCC, Geneva. Nakanishi, N., 2009. Lafarge Aims to make Concrete Making Cleaner. Reuters (accessed 02.02.11.). http://uk.reuters.com/article/2009/04/03/btscenes-uslafarge-environment-intervie-idUKTRE5323W820090403. Naranjo, M., Brownlow, D.T., Garza, A., 2011. CO2 capture and sequestration in the cement industry. Energy Procedia 4, 2716e2723. National Energy Technology Laboratory, 2007. Cost and Performance Baseline for Fossil Energy Plants Volume 1: Bituminous Coal and Natural Gas to Electricity. DOE/NETL-2007/1281. Van Puyvelde, D.R., 2009. CCS opportunities in the Australian Industrial Processes sector. Energy Procedia 1, 109e116. Ridha, F.N., Manovic, V., Macch, A., Anthony, E.J., 2012. The effect of SO2 on CO2 capture by CaO-based pellets prepared with a kaolin derived Al(OH)3 binder. Appl. Energy 92, 415e420. Rodriguez, N., Alonso, M., Grasa, G., Abanades, J.C., 2008a. Heat requirements in a calciner of CaCO3 integrated in a CO2 capture system using CaO. Chem. Eng. J. 138, 148e154. Rodriguez, N., Alonso, M., Grasa, G., Abanades, J.C., 2008b. Process for capturing CO2 arising from the calcination of the CaCO3 used in cement manufacture. Environ. Sci. Technol. 42, 6980e6984. Rodriguez, N., Alonso, M., Abanades, J.C., 2010. Average activity of CaO particles in a calcium looping system. Chem. Eng. J. 156 (2), 388e394. Sarlis, J., Shaw, D., 2008. Canslov Activities and Technology Focus for CO2 Capture. 11th Workshop of the Post-combustion Network, Vienna. Shimizu, T., Hirama, T., Hosoda, H., Kitano, K., Inagaki, M., Tejima, K., 1999. A twin uid-bed reactor for removal of CO2 from combustion processes. Chem. Eng. Res. Des. 77 (A1), 62e68. Taylor, M., Tam, C., Gielen, D., 2006. Energy Efciency and CO2 Emissions from the Global Cement Industry. Workshop on Energy Efciency and CO2 Emission Reduction Potentials and Policies, IEA-WBCSD. WHD Microanalysis Consultants Ltd website, 2011. Clinker: Compositional Parameters. Available at: http://www.understanding-cement.com/parameters.html. Wiemr, T., Berger, R., Hawthorne, C., Abanades, J.C., 2007. Lime enhanced gasication of solid fuels. Fuel 87, 1678e1686. Worrell, E., Martin, N., Price, L., 2000. Potentials for energy efciency improvement in the US cement industry. Energy, 1189e1214. Worrell, E., Price, L., Martin, N., Henrdiks, C., Ozawa Meida, L., 2001. Carbon dioxide emissions from the global cement industry. Annu. Rev. Energy Environ., 303e329. Zeman, F., Lackner, K., 2006. The Zero Emission Kiln. International Cement Review, 55e58.

You might also like