WO2007002882A2 - Regeneration of calcium sulfide to mesoporous calcium carbonate using ionic dispersants and selective reclamation... - Google Patents

Regeneration of calcium sulfide to mesoporous calcium carbonate using ionic dispersants and selective reclamation... Download PDF

Info

Publication number
WO2007002882A2
WO2007002882A2 PCT/US2006/025493 US2006025493W WO2007002882A2 WO 2007002882 A2 WO2007002882 A2 WO 2007002882A2 US 2006025493 W US2006025493 W US 2006025493W WO 2007002882 A2 WO2007002882 A2 WO 2007002882A2
Authority
WO
WIPO (PCT)
Prior art keywords
sorbent
gas
cao
carbon dioxide
reaction
Prior art date
Application number
PCT/US2006/025493
Other languages
French (fr)
Other versions
WO2007002882A3 (en
Inventor
Liang-Shih Fan
Himanshu Gupta
Mahesh V. Iyer
Original Assignee
The Ohio State University
Priority date (The priority date is an assumption and is not a legal conclusion. Google has not performed a legal analysis and makes no representation as to the accuracy of the date listed.)
Filing date
Publication date
Priority claimed from US11/255,099 external-priority patent/US7618606B2/en
Application filed by The Ohio State University filed Critical The Ohio State University
Publication of WO2007002882A2 publication Critical patent/WO2007002882A2/en
Publication of WO2007002882A3 publication Critical patent/WO2007002882A3/en

Links

Classifications

    • BPERFORMING OPERATIONS; TRANSPORTING
    • B01PHYSICAL OR CHEMICAL PROCESSES OR APPARATUS IN GENERAL
    • B01DSEPARATION
    • B01D53/00Separation of gases or vapours; Recovering vapours of volatile solvents from gases; Chemical or biological purification of waste gases, e.g. engine exhaust gases, smoke, fumes, flue gases, aerosols
    • B01D53/34Chemical or biological purification of waste gases
    • B01D53/46Removing components of defined structure
    • B01D53/62Carbon oxides
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01BNON-METALLIC ELEMENTS; COMPOUNDS THEREOF; METALLOIDS OR COMPOUNDS THEREOF NOT COVERED BY SUBCLASS C01C
    • C01B32/00Carbon; Compounds thereof
    • C01B32/60Preparation of carbonates or bicarbonates in general
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01FCOMPOUNDS OF THE METALS BERYLLIUM, MAGNESIUM, ALUMINIUM, CALCIUM, STRONTIUM, BARIUM, RADIUM, THORIUM, OR OF THE RARE-EARTH METALS
    • C01F11/00Compounds of calcium, strontium, or barium
    • C01F11/02Oxides or hydroxides
    • C01F11/04Oxides or hydroxides by thermal decomposition
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01FCOMPOUNDS OF THE METALS BERYLLIUM, MAGNESIUM, ALUMINIUM, CALCIUM, STRONTIUM, BARIUM, RADIUM, THORIUM, OR OF THE RARE-EARTH METALS
    • C01F11/00Compounds of calcium, strontium, or barium
    • C01F11/02Oxides or hydroxides
    • C01F11/04Oxides or hydroxides by thermal decomposition
    • C01F11/06Oxides or hydroxides by thermal decomposition of carbonates
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01FCOMPOUNDS OF THE METALS BERYLLIUM, MAGNESIUM, ALUMINIUM, CALCIUM, STRONTIUM, BARIUM, RADIUM, THORIUM, OR OF THE RARE-EARTH METALS
    • C01F11/00Compounds of calcium, strontium, or barium
    • C01F11/18Carbonates
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01FCOMPOUNDS OF THE METALS BERYLLIUM, MAGNESIUM, ALUMINIUM, CALCIUM, STRONTIUM, BARIUM, RADIUM, THORIUM, OR OF THE RARE-EARTH METALS
    • C01F11/00Compounds of calcium, strontium, or barium
    • C01F11/18Carbonates
    • C01F11/182Preparation of calcium carbonate by carbonation of aqueous solutions and characterised by an additive other than CaCO3-seeds
    • BPERFORMING OPERATIONS; TRANSPORTING
    • B01PHYSICAL OR CHEMICAL PROCESSES OR APPARATUS IN GENERAL
    • B01DSEPARATION
    • B01D2251/00Reactants
    • B01D2251/40Alkaline earth metal or magnesium compounds
    • B01D2251/404Alkaline earth metal or magnesium compounds of calcium
    • BPERFORMING OPERATIONS; TRANSPORTING
    • B01PHYSICAL OR CHEMICAL PROCESSES OR APPARATUS IN GENERAL
    • B01DSEPARATION
    • B01D2251/00Reactants
    • B01D2251/60Inorganic bases or salts
    • B01D2251/602Oxides
    • BPERFORMING OPERATIONS; TRANSPORTING
    • B01PHYSICAL OR CHEMICAL PROCESSES OR APPARATUS IN GENERAL
    • B01DSEPARATION
    • B01D2257/00Components to be removed
    • B01D2257/50Carbon oxides
    • B01D2257/504Carbon dioxide
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01PINDEXING SCHEME RELATING TO STRUCTURAL AND PHYSICAL ASPECTS OF SOLID INORGANIC COMPOUNDS
    • C01P2006/00Physical properties of inorganic compounds
    • C01P2006/12Surface area
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01PINDEXING SCHEME RELATING TO STRUCTURAL AND PHYSICAL ASPECTS OF SOLID INORGANIC COMPOUNDS
    • C01P2006/00Physical properties of inorganic compounds
    • C01P2006/14Pore volume
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01PINDEXING SCHEME RELATING TO STRUCTURAL AND PHYSICAL ASPECTS OF SOLID INORGANIC COMPOUNDS
    • C01P2006/00Physical properties of inorganic compounds
    • C01P2006/16Pore diameter
    • YGENERAL TAGGING OF NEW TECHNOLOGICAL DEVELOPMENTS; GENERAL TAGGING OF CROSS-SECTIONAL TECHNOLOGIES SPANNING OVER SEVERAL SECTIONS OF THE IPC; TECHNICAL SUBJECTS COVERED BY FORMER USPC CROSS-REFERENCE ART COLLECTIONS [XRACs] AND DIGESTS
    • Y02TECHNOLOGIES OR APPLICATIONS FOR MITIGATION OR ADAPTATION AGAINST CLIMATE CHANGE
    • Y02CCAPTURE, STORAGE, SEQUESTRATION OR DISPOSAL OF GREENHOUSE GASES [GHG]
    • Y02C20/00Capture or disposal of greenhouse gases
    • Y02C20/40Capture or disposal of greenhouse gases of CO2
    • YGENERAL TAGGING OF NEW TECHNOLOGICAL DEVELOPMENTS; GENERAL TAGGING OF CROSS-SECTIONAL TECHNOLOGIES SPANNING OVER SEVERAL SECTIONS OF THE IPC; TECHNICAL SUBJECTS COVERED BY FORMER USPC CROSS-REFERENCE ART COLLECTIONS [XRACs] AND DIGESTS
    • Y02TECHNOLOGIES OR APPLICATIONS FOR MITIGATION OR ADAPTATION AGAINST CLIMATE CHANGE
    • Y02PCLIMATE CHANGE MITIGATION TECHNOLOGIES IN THE PRODUCTION OR PROCESSING OF GOODS
    • Y02P20/00Technologies relating to chemical industry
    • Y02P20/151Reduction of greenhouse gas [GHG] emissions, e.g. CO2

Definitions

  • the present invention relates to the application of chemical sorbents for the separation of CO 2 from gas mixtures.
  • the term "supersorbent” shall mean a sorbent as taught in United States Patent No. 5,779,464 entitled “Calcium Carbonate Sorbent and Methods of Making and Using Same", the teachings of which are hereby incorporated by reference.
  • microporous shall mean a pore size distribution of less than 5 nanometers.
  • meoporous shall mean a pore size distribution of from about 5 nanometers to about 20 nanometers.
  • Atmospheric CO 2 concentration has been increasing steadily since the industrial revolution. It has been widely accepted that the while the CO 2 concentration was about 280 ppm before the industrial revolution, it has increased from 315 ppmv in 1959 to 370 ppmv in 2001 [Keeling, CD. and TP. Whorf. 2002. Atmospheric CO 2 records from sites in the SIO air sampling network. In Trends: A Compendium of Data on Global Change.
  • Adsorption systems capture CO 2 on a bed of adsorbent materials such as molecular sieves and activated carbon [Kikkinides, E.S.; Yang, R.T.; Cho, S. H. Concentration and Recovery of CO 2 from flue gas by pressure swing adsorption. Ind. Eng. Chem. Res. 1993, 32, 2714-2720]. CO 2 can also be separated from the other gases by condensing it out at cryogenic temperatures. Polymers, metals such as palladium, and molecular sieves are being evaluated for membrane based separation processes [Reimer, P.; Audus, H.; Smith, A. Carbon Dioxide Capture from Power Stations. IEA Greenhouse R&D Programme, www.ieagreen.org.uk, 2001. ISBN 1 898373 15 9].
  • reaction based processes can be applied to separate CO 2 from gas mixtures.
  • This process is based on a heterogeneous gas- solid non-catalytic carbonation reaction where gaseous CO 2 reacts with solid metal oxide (represented by MO) to yield the metal carbonate (MCO 3 ).
  • MO solid metal oxide
  • MCO 3 metal carbonate
  • the calcination reaction can be represented by:
  • Figure 1 shows the variation in the free energy of the carbonation reaction as a function of temperature for calcium oxide. From the figure, we can see that the carbonation reaction is thermodynamically favored with a decrease in temperature (Gibbs free energy declines with a decrease in temperature). However, at lower temperatures, the carbonation reaction is kinetically slow. In fact, it takes geological time scales for the formation of CaCO 3 by the reaction between CaO and atmospheric CO 2 (at 280-360 ppm) at ambient temperatures. It should also be noted that the carbonation reaction would be favored as long as the free energy is negative. This creates an upper bound of 890 0 C for carbonation to occur under a CO 2 partial pressure of 1 atm.
  • the equilibrium temperature for this reaction is a function of the partial pressure Of CO 2 .
  • a reaction based CO 2 separation process offers many advantages. Under ideal conditions, MEA captures 6Og CO 2 /kg, silica gel adsorbs 13.2g CO 2 /kg and activated carbon adsorbs 88g CO 2 /kg.
  • the sorption capacity of some metal oxides (such as the modified CaO, presented in this study) is about 70Og CO 2 /kg of CaO. This is about an order of magnitude higher than the capacity of adsorbents/solvents used in other CO 2 separation processes and would significantly reduce the size of the reactors and the material handling associated with CO 2 separation.
  • MnCO 3 undergoes a more complex thermal degradation phenomena. MnCO 3 first decomposes to MnO 2 at 300 0 C, which in turn changes to Mn 2 O 3 at 440 0 C. At higher temperatures (-900 0 C), the final thermal decomposition product was identified as Mn 3 O 4 (Shaheen, W. M.; Selim, M. M. Effect of thermal treatment on physicochemical properties of pure and mixed manganese carbonate and basic copper carbonate. Thermochim. Acta. 1998, 322(2), 117-128.).
  • the carbonation extent of Mg(OH) 2 was about 10% between 387-400 0 C and 6% formation between 475-500 0 C (Butt, DP.; Lackner, K.S.; Wendt, C.H.; Conzone, S.D.; Kung, H.; Lu, Y- C; Bremser, J. K. Kinetics of Thermal Dehydroxylation and Carbonation of Magnesium Hydroxide. J. Am. Ceram. Soc.1996, 79(7), 1892-1898). They attributed the low conversions to the formation of a non-porous carbonate product layer. This layer hinders the inward diffusion of CO 2 and the outward diffusion of H 2 O (a product of the carbonation reaction) leading to low conversions.
  • the gas-solid CaO-CO 2 reaction proceeds through two rate-controlling regimes.
  • the first regime involves a rapid, heterogeneous chemical reaction.
  • the reaction slows down due to the formation of an impervious layer of CaCO 3 .
  • This product layer prevents the exposure of unreacted CaO in the particle core to CO 2 for further carbonation.
  • the kinetics of the second regime is governed by the diffusion of ions through the CaCO 3 product layer.
  • the activation energy was estimated to be 21 kcal/mol below 688 K and 43 kcal/mol above it for the product layer diffusion, based on the counter migration of CO 3 2" and O 2" ions through the product layer (Bhatia, S. K.; and Perlmutter, D. D. Effect of the product layer on the kinetics of the CO 2 -Lime Reaction. AIChE J. 1983, 29(1), 79-86).
  • Barker suggested that the CaCO 3 layer is about 22 nm thick and his latter work showed repeated 93% conversion over 30 cycles at 629 0 C on 10 nm CaO particles.
  • cyclical studies conducted at a carbonation temperature of 880 0 C and calcination at 860 0 C led to a drop in conversion from 70% in the first carbonation to 38% in the 7 th carbonation step (Kato, Y.; Harada, N.; Yoshizawa, Y. Kinetic feasibility of a chemical heat pump for heat utilization from high temperature processes. Applied Thermal Engineering. 1999, 19, 239-254).
  • the process described here leads to > 95% conversion due to the application of novel mesoporous CaO sorbents for CO 2 capture and maintains their reactivity over repeated cycles of carbonation and calcination.
  • Water gas a mixture of CO, CO 2 , H 2 O and H 2 , is formed by the gasification of coal by sub-stoichiometric air and/or steam. Irrespective of the initial concentration of these four gases, the reversible water gas shift (WGS) reaction gets initiated until the exact ratio of the concentration of these gases reaches a particular equilibrium constant KWGS that is a function of temperature.
  • the WGS reaction and its equilibrium constant can be written as:
  • dolomite does not entirely contain calcium based material.
  • dolomite comprises of nearly 50wt.% calcium, which participates in the reaction to some extent, and the remaining portion of the sorbent (mainly magnesium oxide) stays unreacted. Further, they attribute the incomplete conversions of the calcium material to the concept of pore filling and pluggage at the pore-mouths of these sorbent particles by CaCO 3 product layer, preventing the access of CO 2 in the gas to unreacted CaO surface at the pore interiors.
  • Calcination in a pure CO 2 stream will result in higher operating temperatures due to the thermodynamic limitations of the calcination reaction in presence of the CO 2 product. Higher temperatures and the presence of CO 2 during calcination would cause the sorbent to sinter. This is in agreement with the results of multiple carbonation-calcination cycle tests for dolomite by Harrison and co-workers (Lopez Ortiz and Harrison, 2001) in pure CO 2 stream (800-950 0 C). They observed a decrease in "calcium" conversion from 83 % in the 1 st cycle to about 69 % in the 10 th cycle itself.
  • a preferred method for separating carbon dioxide from a flow of gas comprising carbon dioxide comprises the steps of: (1) directing the flow of gas to a gas-solid contact reactor, the gas-solid contact reactor contains at least one sorbent comprising at least one metal oxide; (2) reacting the carbon dioxide with the at least one sorbent so as to remove the carbon dioxide from said flow of gas, thereby converting the at least one sorbent into spent sorbent; (3) calcining the spent sorbent so as to liberate the carbon dioxide from the spent sorbent, thereby regenerating the sorbent; and (4) repeating the aforementioned steps.
  • the at least one metal oxide is selected from the group consisting of: ZnO, MgO, MnO 2 , NiO, CuO, PbO, and CaO.
  • the spent sorbent is a metal carbonate. It is preferred that the sorbent has a sorption capacity of at least about
  • the sorbent has a sorption capacity of at least about 300 grams of carbon dioxide per kilogram of sorbent. Irrespective of the sorption capacity of the sorbent, it is preferred that the sorbent has substantially the same sorption capacity after calcining as the sorbent had prior to adsorbing the carbon dioxide.
  • any calcination method may be employed, it is preferred that the calcining is performed under at least partial vacuum. It is also preferred that the calcining is performed by steam.
  • the present invention includes facilities practicing the aforementioned method.
  • a method for separating carbon dioxide from a flow of gas comprising carbon dioxide of the present invention comprises the steps of: (1 ) directing the flow of gas to a first gas-solid contact reactor, the first gas-solid contact reactor containing at least one sorbent, the sorbent comprising at least one metal oxide; (2) reacting the carbon dioxide in the flow of gas on the sorbent in the first gas-solid contact reactor so as to remove the carbon dioxide from the flow of gas; (3) directing the flow of gas to a second gas-solid contact reactor when the sorbent in the first gas-solid contact reactor is spent thereby forming spent sorbent, the second gas-solid contact reactor containing at least one sorbent, the sorbent comprising at least one metal oxide; (4) reacting the carbon dioxide in the flow of gas on the sorbent in the second gas-solid contact reactor so as to remove the carbon dioxide from the flow of gas; (5) calcining the spent sorbent from the first gas-solid contact reactor so as to generate carbon dioxide and to regenerate the sorbent;
  • any calcination method may be employed, it is preferred that the calcining is performed under at least partial vacuum. It is also preferred that the calcining is performed by steam. This applies to both gas-solid contact reactors.
  • the at least one metal oxide is selected from the group consisting of: ZnO, MgO, MnO 2 ,
  • the sorbent has a sorption capacity of at least about
  • the sorbent has a sorption capacity of at least about 300 grams of carbon dioxide per kilogram of sorbent. Irrespective of the sorption capacity of the sorbent, it is preferred that the sorbent has substantially the same sorption capacity after calcining as the sorbent had prior to adsorbing the carbon dioxide.
  • the present invention also includes facilities practicing the aforementioned method
  • a method for regenerating a spent sorbent for carbon dioxide of the present invention comprises the steps of: (1) providing a spent sorbent, the spent sorbent comprising metal carbonate; and (2) calcining the spent sorbent so as to liberate carbon dioxide gas and so as to regenerate the spent sorbent thereby forming a sorbent comprising a metal oxide.
  • the spent sorbent is calcium carbonate. It is further preferred that the metal oxide is calcium oxide.
  • the sorbent has substantially the same sorption capacity after calcining as the sorbent had prior to adsorbing the carbon dioxide.
  • any calcination method may be employed, it is preferred that the calcining is performed under at least partial vacuum. It is also preferred that the calcining is performed by steam. This applies to both gas-solid contact reactors.
  • the present invention includes facilities practicing the aforementioned method.
  • a method for producing a sorbent of the present invention comprises the steps of: (1 ) obtaining a structurally altered high surface area calcium carbonate having a surface area of at least 25.0 m 2 /g, a pore volume of at least 0.05 cm 3 /g, and a mesoporous pore size distribution; and (2) calcining the structurally altered high surface area calcium carbonate so as to produce a sorbent having a surface area of less than 22 m 2 /g, a pore volume of at least 0.005 cm 3 /g, and a mesoporous pore size distribution.
  • any calcination method may be employed, it is preferred that the calcining is performed under at least partial vacuum.
  • the present invention includes sorbents made according to the aforementioned method.
  • a sorbent according to the present invention comprising calcium oxide having a surface area of at least 12.0 m 2 /g and a pore volume of at least 0.015 cm 3 /g, the calcium carbonate sorbent having sorption capacity of at least about 70 grams of carbon dioxide per kilogram of sorbent.
  • Figure 1 depicts the Gibbs Free Energy diagram for the carbonation reaction, CaC ⁇ 3 ⁇ CaO + CO 2 , as a function of temperature.
  • Figure 2 illustrates the performance of calcium oxide for the carbonation reaction.
  • Figure 3 compares the XRD diffractograms of CaO derived from various precursors.
  • Figure 4 is a schematic diagram of a carbonator reactor for the synthesis of precipitated calcium carbonate.
  • Figure 5 shows the change in the pH of the slurry as a function of
  • Precipitated Calcium Carbonate (500 mL water, 0.0575% N40V ® dispersant, 4 scfh CO 2 ).
  • Figure 7 compares the pore size distribution of four CaO precursors.
  • Figure 8 compares the conversion of four CaO sorbents under pure CO 2 at 650 0 C.
  • Figure 9 illustrates the effect of temperature on the carbonation of
  • Figure 10 illustrates the carbonation-calcination cycles on Aldrich
  • FIG. 11 shows extended carbonation-calcination cycles on precipitated calcium carbonate (PCC) powder at 700 0 C.
  • Figure 12 compares the effect of initial surface area of PCC-CaO to its reactivity towards the carbonation reaction at 700 0 C.
  • Figure 13 depicts the effect of vacuum calcination on the reactivity of
  • Figure 14 provides a flow sheet depicting the integration of the current process in the overall coal-gasifier electric production facility.
  • Figure 15 illustrates thermodynamic data for predicting the temperature zones for hydration and carbonation of CaO.
  • Figure 16 illustrates thermodynamic data for predicting the equilibrium
  • Figure 17 shows a modified reactor set-up with steam generating unit for investigating WGS and carbonation reactions.
  • Figure 18 illustrates the set-up for combined vacuum/sweep gas calcination experiments allowing the use of larger sorbent samples.
  • Figure 19 is a pore size distribution of the HTS and LTS obtained from
  • Figure 20 shows the pore size distribution of various calcium oxide precursors.
  • Figure 22 shows the extent of reaction equilibrium as a function of temperature for the WGS reaction.
  • Figure 26 depicts a typical steam generation scenario and use.
  • Figure 27 depicts one implementation of one embodiment of the present invention.
  • Figure 28 depicts one implementation of one embodiment of the present invention.
  • Figure 29 depicts one implementation of one embodiment of the present invention.
  • Figure 30 depicts one implementation of one embodiment of the present invention.
  • Figure 31 depicts one implementation of one embodiment of the present invention.
  • Figure 32 depicts one implementation of one embodiment of the present invention.
  • Figure 33 depicts one implementation of one embodiment of the present invention.
  • Figure 34 depicts one implementation of one embodiment of the present invention.
  • Figure 35 depicts one implementation of one embodiment of the present invention.
  • Figure 36 depicts one implementation of one embodiment of the present invention.
  • Figure 37 depicts one implementation of one embodiment of the present invention.
  • Figure 38 illustrates the calcination and subsequent sulfation of the precipitated calcium carbonate.
  • Figure 39 provides an XRD plot obtained after drying the product.
  • Figure 40 provides an XRD plot of the solid product after 10 minutes of carbonation.
  • Figure 41 provides an XRD plot of the solid product after 60 minutes of carbonation.
  • Figure 42 compares the pore size distribution for two CaCO 3 products.
  • Sorbent Reactivity Testing and Structural Analysis [0086] The reactivity testing of CaO sorbents for carbonation was carried out in a Perkin Elmer Thermogravimetric Analyzer (TGA-7) apparatus. The balance can accurately measure up to 1 microgram. A small sample of the sorbent (5-20 mg) is placed in a quartz boat. The weight of the sample was recorded every second. The structural properties of CaO sorbents and their precursors were tested in a NOVA 2200 analyzer (Quantachrome Company). The BET surface area, pore volume, and pore size distribution were measured at -196 0 C using nitrogen as the adsorbent.
  • Metal oxides such as ZnO, MgO, CuO, MnO 2 , NiO, PbO and CaO that undergo the CCR scheme in the 800-200 0 C temperature range were analyzed for their reactivity in a TGA.
  • a powdered sample of these oxides was placed in a quartz pan and pure CO 2 was passed over the sample metal oxide. The temperature was then slowly raised and the weight of the sample was continuously monitored. An increase in the weight of the sample is an indication of the formation of metal carbonate.
  • Figure 2 provides experimental data for the carbonation of lime (Ca(OH) 2 ) under flowing pure CO 2 gas. With an increase in temperature, the weight of the sample increases till the temperature reaches about 890 0 C.
  • Calcination which is thermodynamically favored above 890 0 C at 1 atm CO 2 partial pressure, causes a rapid decrease in weight until the sorbent converts completely to CaO.
  • the weight starts to increase again and the process is repeated once more.
  • the data also shows recyclability of the sorbent.
  • CaO was identified as a viable candidate for the carbonation- calcination reactions.
  • precursors can be calcined to obtain the CaO sorbents necessary for the carbonation reaction.
  • Common and economical precursors include calcium carbonate, calcium hydroxide and dolomite.
  • the other important source of CaO is via the calcination of synthesized high surface area precipitated calcium carbonate.
  • XRD patterns were obtained on all the CaO sorbents. Figure 3 depicts these d iff ractog rams (a. Calcined Aldrich-CaO; b.
  • a K-type thermocouple 4Oh inserted in the slurry continuously records the slurry temperature.
  • a pH probe 40b monitors the pH of the slurry as the reaction medium changes from a basic to an acidic solution as the reaction proceeds. First, 500 ml of distilled water is poured into the carbonator, followed by the addition of 0.0575g of N40V ® . 12.8g of Ca(OH) 2 is added to the solution to provide a loading of 2.56% by weight. This corresponds to a concentration of 16-sat (concentration of Ca(OH) 2 is 16 times its saturation solubility limit).
  • CaCO 3 has a much lower solubility in water ( ⁇ 0.0012 g/10Og water) compared to Ca(OH) 2 and thus precipitates out.
  • Ca 2+ ions get depleted, but are continuously replenished by the suspended Ca(OH) 2 .
  • the pH remains 12.
  • Ca(OH) 2 ultimately gets depleted and the concentration of Ca 2+ ions cannot be maintained at its solubility limit.
  • continued dissolution of CO 2 gas leads to the accumulation of H + ions causing the solution to become acidic.
  • the pH settles at about 6.0, corresponding to equilibrium solubility of CO 2 in water at ambient temperature.
  • Precipitated calcium carbonate can be obtained by the reaction between carbonate and calcium ions in solution. It is known that the CaCO3 nuclei that precipitate out have positive surface charge on them that prevent agglomeration (Agnihotri, R.; Chauk, S.; Mahuli, S.; Fan, L.-S. Influence of Surface Modifiers on Structure of Precipitated Calcium Carbonate. Ind. Eng. Chem. Res. 1999, 38, 2283- 2291). The resulting structure is also microporous in nature. However, the structural properties of the synthesized PCC can be altered by the use of negatively charged dispersants that neutralize the surface charges.
  • Table 1 Morphological properties of PCC as a function of N40V ® :Ca(OH) 2 loading ratio (500 ml water, 0.0575% N40V ® dispersant, 4 scfh CO 2 ).
  • CaO sorbents were synthesized by calcining various CaO precursors such as Linwood calcium carbonate (LC), dolomite (DL), Linwood calcium hydroxide (LH), and precipitated calcium carbonate (PCC).
  • CaO precursors such as Linwood calcium carbonate (LC), dolomite (DL), Linwood calcium hydroxide (LH), and precipitated calcium carbonate (PCC).
  • LC-CaO Linwood calcium carbonate
  • DL dolomite
  • LH Linwood calcium hydroxide
  • PCC precipitated calcium carbonate
  • the oxides derived from these sources are termed as LC-CaO, FCD-CaO (for fully calcined dolomite-CaO), LH-CaO, and PCC-CaO, respectively.
  • the procedure involved heating the precursor in flowing nitrogen beyond the calcination temperature (800- 950 0 C) for an hour followed by its storage in a desiccator.
  • Structural properties such as surface area (SA) and pore volume (PV) of these chemicals are listed in Table 2 and their pore size distributions are shown in Figure 7.
  • SA surface area
  • PV pore volume
  • Table 2 Morphological properties (surface area and pore volume) of various CaO sorbents and their precursors.
  • Cyclic calcination and carbonation [0093] One of the possible hurdles in the utilization of metal oxides for the carbonation and calcination reaction scheme is its vulnerability to sintering due to the thermal cycling imposed by the cyclical nature of these reactions. Cyclical studies were carried out to quantify any loss in reactivity of these sorbents upon multiple cycles. The temperature chosen for cyclical studies was 700 0 C. This temperature is sufficient to achieve carbonation in the presence of pure CO 2 , and also to calcine the CaCO 3 so formed after the gas is switched from CO2 to N 2 . A variety of precursors were first calcined in nitrogen at 700 0 C. The gas was then switched to pure CO 2 and the weight gain continuously tracked.
  • Vacuum calcination leads to the formation of a metastable-nanocrystalline calcia structure while calcination in helium atmosphere lead to a stable microcrystalline calcia structure (Dash, S., Kamruddin, M., Ajikumar, P.K., Tyagi, A.K., and Raj, B., "Nanocrystalline and metastable phase formation in vacuum thermal decomposition of calcium carbonate.” Thermochimica acta, 2000, 363, 129-135).
  • Beruto et al., [1980] estimated the surface area and pore volume of limestone based CaO to be about 78-89 m 2 /g and 0.269 ml/g respectively.
  • Table 3 Structural properties of Calcium based sorbents undergoing vacuum calcination at 750 0 C and carbonation at 700 0 C.
  • a variety of chemical processes known to generate syngas include:
  • the flow sheet shown in Figure 14 integrates the Calcium-based reactive separation process under development in this project with a coal gasifier based electric power/chemical synthesis process plant 140.
  • the main coal gasifier 140a consists of a high pressure and high temperature unit that allows contact between coal 140b, steam 14Oe and air/pure oxygen 14Oy in a variety of schemes.
  • Boiler feed water 14Od is preheated by passing it through gasifier 140a prior to steam tubine 140c. Waste from the gasifier is collected as slag 14Oz.
  • Typical fuel gas compositions from various known coal gasifiers are shown in Table 4.
  • the reacted CaCO 3 particles are captured using a high temperature solids separator 14Oh (e.g., a candle filter or a high temperature ESP) and separated fuel gas stream.
  • the spent solids are now sent to a rotary calciner 140k to thermally decompose the CaCO 3 14Oj back to CaO 14Of and pure CO 2 140m.
  • the high purity CO 2 gas can now be economically compressed 1401, cooled, liquefied and transported for its safe sequestration 140m.
  • the rotary calciner allows the calcium particles to remain segregated, which is crucial in maintaining a sorbent structure characterized by a higher porosity.
  • the calcination of the sorbent can also be effected under sub-atmospheric conditions that allow the removal of CO 2 as soon as it is formed from the vicinity of the calcining sorbent, thereby aiding further calcination.
  • This vacuum can be created by means of ejector systems that are widely used in maintaining vacuum in large vacuum distillation units (VDU) in the petroleum refining industry. Lock and hopper combinations and appropriate seals ensure that the sorbent can be effectively separated from the CO 2 stream and re-entrained in the fuel gas duct.
  • the hydrogen enriched fuel gas 14Oi can now be used to generate electric power in a fuel cell 14On or used to make fuels and chemicals 14Oq without any low temperature clean up.
  • the fuel cell may receive a supply of air 14Op and discharge steam 140o.
  • the hydrogen enriched fuel gas may be sent to gas turbine 14Or used to drive generator 14Ot to produce electricity and air compressor 140s to produce a stream of compressed air.
  • the stream of compressed air may be sent to air separator 14Ox to produce the air/oxygen of 14Oy.
  • the discharge from gas turbine 14Ot may be sent through heat exchanger 14Ou prior to being discharged at stack 14Ov.
  • the absorbed heat may be collected by steam turbine 14Ow to produce additional electricity.
  • CaO calcium oxide
  • Table 4 Typical fuel gas compositions obtained from different gasifiers. (Stultz and Kitto, 1992)
  • FIG. 15 (b) shows the typical equilibrium CO 2 partial pressures (PCO 2 ) as a function of temperature. From the data in Table 4, it can be inferred that the typical PCO2 in the gasifiers ranges from 0.4-4.3 atm for entrained flow (slurry) and entrained flow (dry) gasifier systems respectively.
  • the high and low temperature water gas shift (WGS) reaction catalysts were procured from S ⁇ d-Chemie Inc., Louisville, KY.
  • the high temperature shift (HTS) catalysts comprises of iron (III) oxide supported on chromium oxide.
  • Precipitated calcium carbonate (PCC) was synthesized by bubbling CO2 through a slurry of hydrated lime.
  • the neutralization of the positive surface charges on the CaCO 3 nuclei by negatively charged N40V ® molecules forms CaCO 3 particles characterized by a higher surface area/pore volume and a predominantly mesoporous structure. Details of this synthesis procedure have been reported elsewhere (Agnihotri et al., 1999). Hydrated lime from a naturally occurring limestone (Linwood Hydrate, LH) and a naturally occurring limestone (Linwood Carbonate, LC) was obtained from Linwood Mining and Minerals Co.
  • the sorbents and catalyst were analyzed to determine their morphologies using a BET analyzer.
  • the BET surface areas, pore volumes, and pore size distributions of the catalysts and sorbents were measured at -196 0 C using nitrogen as the adsorbent in a Nova 2200 Quantachrome BET analyzer. Special care was taken to ensure that all samples were vacuum degassed at 250 0 C for 5 hours prior to BET analysis.
  • a reactor setup was designed, underwent several iterations and was assembled to carry out water gas shift reactions in the presence of CaO and catalyst.
  • the reactor design assembly used to carry out these experiments is shown in Figure 17.
  • This setup enables us to carry out both the water gas shift reaction in the presence of CaO as well as the regeneration of the sorbent in flowing gas such as nitrogen and/or steam.
  • the setup 170 consists of a tube furnace 17Op, a steel tube reactor 170a, a steam generating unit 170c, a set of gas analyzers for the online monitoring of CO and CO 2 concentrations 17On, a condenser 170m to remove water from the exit gas stream and a high pressure water syringe pump 170b.
  • All the reactant gases (H 2 , CO, CO 2 , and N 2 ) are metered using modified variable area flowmeters 17Oe - h respectively.
  • the syringe pump is used to supply very accurate flow-rates of water into the heated zone of the steam- generating unit in the 0.01-0.5 ml/min range.
  • the steam generator is also packed with quartz wool 17Od in order to distribute the water drops as they enter into the heating zone.
  • the packing is utilized in order to provide greater surface for water evaporation and to dampen out fluctuations in steam formation.
  • the main problem with a fluctuating steam supply is that the gas analyzers used to measure the exit CO and CO 2 concentrations are sensitive to gas flow rates. Even though the steam is being condensed out before the gas is sent into the analyzers, surges in the steam supply still affect the overall gas flow rate, causing the CO and CO 2 readings to fluctuate.
  • the packing ultimately ensures a more continuous and constant overall gas flow rate into the main reactor and into the analyzers.
  • Thermocouple 170k is used to monitor the temperature inside reactor 170a. Any extra gas inlets of reactor 170a are blocked 170I.
  • a steel tube reactor is used to hold the Ca-based sorbent and catalyst, and is kept heated using a tube furnace.
  • the sorbent loading unit of the reactor is detachable which enables easy removal and loading of the sorbent and therefore minimizes the sorbent loading time between runs. Also, the sorbent can be changed without having to cool down the entire reactor.
  • the gas mixture 17Oj entering the reactor is preheated to the reaction temperature before contacting the sorbent/ catalyst particles.
  • the gases exiting the reactor first flow through a condenser in order to separate out the moisture and then to a set of gas analyzers.
  • Sub Atmospheric Calcination Reactor Setup [0105] Once the Calcium based sorbent has reacted with the CO 2 being produced by the WGSR, the sorbent has to be regenerated for further use in subsequent cycles. During the regeneration of the sorbent, carbon dioxide is released from the sorbent. In order to minimize the necessity for further treatment of this released CO 2 before sending it to sequestration sites, it is necessary to regenerate the sorbent such that a pure stream of CO 2 is released. Vacuum calcination provides one method for ensuring that concentrated streams of CO2 are release in the regeneration phase. The detailed setup is shown in Figure 18. This setup 180 was assembled to handle the regeneration of large quantities of sorbent ( ⁇ 10-20g per batch).
  • the setup includes an alumina tube reactor 180b, which would hold the sorbent samples in a split tube furnace 180c that provides the heat necessary to calcine the sorbent 18Od, two Non Dispersive Infra Red (NDIR) analyzers 180k -I to monitor the CO 2 concentration (ranges 0-2500ppm and 0-20%) and two vacuum pumps 18Of and 18Oi. 10g of sorbent yields about 2.4L of CO 2 at atmospheric pressure and temperature over the entire decomposition process. This gas needs to be diluted with air in order to ensure that the CO 2 concentration lies in the detection range of the CO 2 analyzers.
  • NDIR Non Dispersive Infra Red
  • Vacuum Pump 18Of is a dry vacuum pump procured from BOC Edwards capable of achieving vacuum levels as low as 50mtorr and gas flowrates of 6m 3 /hr.
  • the CO2 analyzers have their own inbuilt pumps and are capable of drawing up to 2LPM from the header for online CO2 analysis.
  • the second pump 18Oi is a smaller dry pump and is put in place to ensure that there is no pressure buildup in the !4" lines connecting the vacuum pump to the analyzers. Pump 18Oi discharges to vent 18Oj.
  • the temperature of the furnace is controlled with a thermocouple inserted into the central zone of the furnace.
  • the temperature of the reactor was also monitored using a second thermocouple inserted into the center of the alumina tube.
  • the setup is also capable of combining vacuum calcination with flow of sweep gas 180a. As it may not be feasible to supply very low vacuum levels for the calcination of the sorbent in industrial settings, it may be necessary to study the calcination process in combination with the addition of various sweep gases such as N2/ steam. Pressure gauges 18Oe, h and volumetric flow meters are included to monitor the vacuum pressure in the reactor, the pressure in the VA" lines and the flows of the sweep gases into the calciner and the flow of the air 18Og used in the dilution of the exhaust CO 2 before sending it to the analyzers. The analyzers are also connected to a data acquisition system 180m that can record analyzer readings every second. RESULTS AND DISCUSSIONS
  • the low temperature shift (LTS) catalyst has a BET surface area of 85 m /g and a total pore volume of about 0.3 cc/g. The majority of the pores were found to occur around 180 A as evident from the maximum in its pore size distribution plot shown in Figure 19.
  • the low temperature shift (LTS) catalyst has a BET surface
  • the SA/PV of the calcium hydroxide (LH) fines increase as can be observed for the LH sample.
  • the porosity is maximized in the microporous range (30-50 A range).
  • the SA/PV of the morphologically altered PCC are much higher. Further, most of the porosity lies in the 100-300 A range.
  • Table 5 Morphological properties of the natural and synthesized CaO precursors and the HTS catalyst obtained from BET analyses.
  • WGSR Water gas shift reaction
  • the HTS catalyst was tested for its catalyst activity towards the WGS reaction between 500-700 0 C. Blank runs (without any sorbent) were performed in a reaction mixture comprising of 3% CO and 9% H2O, the balance being 5.0 grade N2. The total gas flow-rate was maintained at about 1.5 slpm and the steam/CO ratio was set at ⁇ 3. Typically about 0.5 grams of the HTS catalyst was loaded in the reactor prior to each run. The catalyst activity increases monotonically with increasing reaction temperature. This is evident from Figure 21 below. The CO conversion increases from 24.3% at 500 0 C to 69.3% at 550 0 C. It finally reaches about 80% at 600 0 C. Beyond 600 0 C the conversion does not change much but remains steady at -78% at 700 0 C. This might be due to the equilibrium limitations governing the WGS reaction scheme is depicted in eqn (8) below:
  • thermodynamic equilibrium constant (K) for any temperature for this reaction was computed using the software "HSC Chemistry v 5.0" (Outokumpu Research Oy, Finland).
  • the observed ratio was computed from the experimental data by obtaining the ratio of the partial pressures of the products and the reactants as per the eqn (9) below:
  • FIG. 22 illustrates the effect of temperature on the ratio of partial pressures (Kobs) obtained from the experimental data. This is compared with the thermodynamic equilibrium values (K eq ). From the figure it is evident that we are operating in the region that is below the thermodynamic equilibrium. At 500 °C the Ko b s is 0.028 while the corresponding K eq is 4.77. K eq monotonically decreases with increasing temperature. In contrast, Ko bs increases with temperature for our operating conditions. Thus, at 600 0 C the Ko bs increases to 1.4 while the K eq moves down to 2.5. This trend continues and it is clearly evident from the figure that the system moves closer to equilibrium as we progressively increase the temperature from 500 to 700 0 C. Combined Carbonation and Water gas shift reaction: Sorbent Testing and Analyses
  • the termination of the calcination was ensured by monitoring the CO2 released using a CO 2 analyzer.
  • the reaction temperature was again lowered to 600 0 C and the sorbent-catalyst system was subjected to the reaction mixture for a second reaction cycle.
  • the 2 nd cycle CO breakthrough curve is also depicted in Figure 23. It is evident from the figure that the CO conversion is not as superior as in the 1 st cycle.
  • the CO conversion monotonically decreases to about 90% in 110 seconds, 80% in 240 seconds and gradually to about 50%. It is interesting to note that at the end of the breakthrough the sorbent-free catalytic CO conversion of 81% is not achievable. This could be attributed to the loss in the catalytic activity after the first regeneration cycle.
  • PCC and LH sorbent-catalyst systems The curves are for the 1 st reaction cycle.
  • the CO conversion at any given time for PCC-CaO is always higher than that of LH-CaO.
  • the PCC system gives almost 100% conversion for first 240 seconds (4 min) while the LH sorbent system sustains this conversion only in the initial few seconds. Subsequently, the PCC system gives about 90% CO conversion at 1000 seconds (16.5 min) followed by 85% in 1600 seconds (27 min). In contrast, the LH system gradually gives about 90% CO conversion at 900 seconds (15 min) and followed by 85% in 1200 seconds (20 min). Both the sorbent systems gradually achieve their maximum loading capacity with time and finally at around 2500-3000 seconds they reach their breakthrough loading.
  • Figure 27 illustrates one embodiment of the present invention providing
  • Figure 28 illustrates a second embodiment of the present invention providing 1 MWe total capacity.
  • Figure 29 illustrates another embodiment of the present invention providing 1.33 MWe total capacity.
  • Figure 30 illustrates yet another embodiment of the present invention providing 1.33 MWe total capacity
  • Figure 31 illustrates an alternative embodiment of the present invention providing 1.54 MWe total capacity.
  • Figure 32 illustrates yet another alternative embodiment of the present invention providing 1.07 MWe total capacity.
  • Figure 33 illustrates an alternative embodiment of the present invention providing 1 MWe total capacity.
  • Figure 34 illustrates an alternative embodiment of the present invention providing 1 MWe total capacity.
  • Figure 35 illustrates yet another embodiment of the present invention providing 1.54 MWe total capacity.
  • Figure 36 illustrates an alternative embodiment'of the present invention providing 1 MWe total capacity at 80% CO 2 capture.
  • Figure 37 illustrates another embodiment of the present invention providing 300 MWe total capacity at 90 CO 2 capture.
  • CBS for the control of acid gases such as H2S, CO2 and S02 from combustion and gasification of fossil fuels such as coal, oil and natural gas
  • acid gases such as H2S, CO2 and S02
  • combustion and gasification of fossil fuels such as coal
  • oil and natural gas are hampered by the incomplete reactivity of calcium towards various gas-solid reactions such as sulfation to form CaSO 4 , sulfidation to form CaS, and carbonation to form CaCO3.
  • High conversion of CBS is essential in lowering the levelized operating cost of the concerned process.
  • One of the methods to enhance reactivity of these CBS is to recycle the unreacted sorbent back into the flue and fuel gas mixtures.
  • This process involves the extraction of the unreacted calcium (in the form of CaO and Ca(OH)2) from the solid mixture by an aqueous sugar solution that forms complexes with calcium.
  • Other chemicals that can achieve this complexation include OH-containing groups such as phenols, glycerol, glycols, EDTA, etc.
  • This complexation leads to the selective removal of only unreacted calcium from 1 the solid mixture by forming water-soluble complexes. The slurry is then filtered and the residue now contains fly ash and calcium sulfate (which signifies 100% conversion of calcium). The calcium escapes with the filtrate.
  • the solution is taken into another vessel where it is subjected to carbonation in which gaseous C02 is injected into the sugar solution containing calcium in the form of chemical complexes.
  • the carbonation reaction leads to the formation of calcium carbonate crystals that can be removed by filtration.
  • the CaC03 is now dried and injected for further SO2 capture in the FSI mode.
  • the drop in weight of the sorbent, as shown in Figure 38 to 56% in the calcination phase indicates the formation of pure CaCO3 only. This also proves that other materials, such as fly ash and CaSO4 do not leach out into the sugar solution and mix with the precipitated CaCO3.
  • a CaO based CO2 separation process is under development and patent protection at OSU. It involves the injection of CaO fines into flue/fuel gas mixtures to reactively separate CO2 in the form of CaCO3. These particles are physically removed from the gas duct and separately calcined to release pure C02 stream and CaO, which are used in the next carbonation phase. This cyclic carbonation and calcination reaction (CCR) process leads to CO2 separation.
  • CCR cyclic carbonation and calcination reaction
  • H 2 S and SO 2 attacks the CBS forming CaS and CaSO4 respectively, which reduces the sorbent capacity for C02 capture.
  • the complexation-based process can be used to obviate the challenge that H2S and SO2 pose.
  • the mixture containing the sorbents is treated with complexing agent solution (such as sugar solution) to extract out unreacted CaO as explained above.
  • complexing agent solution such as sugar solution
  • the Calcium carbonate so formed is then sent to the catchier to convert it to CaO. This fresh and pure CaO is then re-introduced into the CCR scheme.
  • This process provides a stream of H 2 S as well as CaCO 3 that can be recycled back for additional H 2 S removal in a subsequent cycle.
  • This leads to a loss in sorbent reactivity after a few cycles of sulfidation regeneration reaction.
  • the sulfidation extent of CaCO 3 dropped from 69% in the first cycle to 13% at the end of 21 cycles (Keaims et al., 1974).
  • Another process involves the high temperature oxidation (at 920 ° C) of CaS to form CaSO 4 , which is more amenable to land filling, as it is a stable and relatively unreactive compound (Qiu, et al., 1999).
  • the use of a regenerative calcium based desulfurization process has additional advantages. For example, the use of calcium over multiple cycles drastically reduces the requirement of fresh calcium sorbent, as a majority of it is recycled. A related benefit is that the quantity of byproducts that require additional management such as land filling is also reduced.
  • Nishev and Pelovski (1993) detailed the effect of reaction temperature and PCO2 on the kinetics of carbonation of CaS.
  • An increase in P C o 2 increases the rate of the reaction as well as the conversion.
  • a higher P C o 2 leads to faster nucleation that leads to a less porous CaCO 3 sorbent.
  • Internal diffusion control dominated the progress of the reaction as evidenced by the activation energy values in the 17-27 kJ/mol. It was also shown that a higher Pco 2 lead to a higher value of activation energy and pre-exponential factor.
  • Brooks and Lynn investigated the recovery of calcium carbonate and hydrogen sulfide from spent calcium sorbents.
  • MDEA methyldiethanolamine
  • This aqueous phase carbonation provides a method to completely regenerate calcium- based sorbents, which lose reactivity towards high temperature reactions that are carried out in a cyclic fashion. Further it is our aim to study the morphology of the CaCO 3 that is formed for its surface area and pore volume characteristics, which are crucial for sorbent reactivity. A prior study has already indicated the efficacy of OSlTs patented mesoporous sorbent towards the extent of sulfidation compared to naturally occurring limestone (Chauk et al., 2000).
  • the main aim of this portion of the project is to recreate the mesoporous PCC structure starting from CaS (pure and CaS resulting from sulfidation of CaO/CaCO 3 ) instead of the conventional Ca(OH) 2 starting material.

Landscapes

  • Chemical & Material Sciences (AREA)
  • Organic Chemistry (AREA)
  • Inorganic Chemistry (AREA)
  • Life Sciences & Earth Sciences (AREA)
  • Geology (AREA)
  • Thermal Sciences (AREA)
  • Engineering & Computer Science (AREA)
  • Physics & Mathematics (AREA)
  • Environmental & Geological Engineering (AREA)
  • Analytical Chemistry (AREA)
  • General Chemical & Material Sciences (AREA)
  • Oil, Petroleum & Natural Gas (AREA)
  • Chemical Kinetics & Catalysis (AREA)
  • Biomedical Technology (AREA)
  • Health & Medical Sciences (AREA)
  • Solid-Sorbent Or Filter-Aiding Compositions (AREA)
  • Treating Waste Gases (AREA)

Abstract

A reaction-based process has been developed for the selective removal of carbon dioxide (CO2) from a multicomponent gas mixture to provide a gaseous stream depleted in CO2 compared to the inlet CO2 concentration in the stream. The proposed process effects the separation of CO2 from a mixture of gases (such as flue gas/fuel gas) by its reaction with metal oxides (such as calcium oxide). The Calcium based Reaction Separation for CO2 (CaRS-CO2) process consists of contacting a CO2 laden gas with calcium oxide (CaO) in a reactor such that CaO captures the CO2 by the formation of calcium carbonate (CaCOa). Once 'spent', CaCO3 is regenerated by its calcination leading to the formation of fresh CaO sorbent and the evolution of a concentrated stream of CO2. The 'regenerated' CaO is then recycled for the further capture of more CO2. This carbonation-calcination cycle forms the basis of the CaRS-CO2 process. This process also identifies the application of a mesoporous CaCO3 structure, developed by a process detailed elsewhere, that attains >90% conversion over multiple carbonation and calcination cycles. Lastly, thermal regeneration (calcination) under vacuum provided a better sorbent structure that maintained reproducible reactivity levels over multiple cycles.

Description

REGENERATION OF CALCIUM SULFIDE TO MESOPOROUS CALCIUM
CARBONATE USING IONIC DISPERSANTS AND SELECTIVE RECLAMATION OF
UNREACTED CALCIUM FROM CALCIUM-CONTAINING SOLD MIXTURES TO
MAXIMIZE CALCIUM CONVERSION AND PREVENT RECYCLING OF INERTS
Inventors: Liang-Shih Fan Himanshu Gupta Mahesh Iyer
RELATED APPLICATION DATA
[0001] This application claims the priority benefit of U.S. Patent Application
Serial No. 60/694,702, filed June 28, 2005, which is hereby incorporated herein by reference. This application also claims the priority benefit of U.S. Patent Application Serial No. 11/255,099, filed October 20, 2005, which is hereby incorporated herein by reference.
TECHNICAL FIELD OF THE INVENTION [0002] The present invention relates to the application of chemical sorbents for the separation of CO2 from gas mixtures.
BACKGROUND OF THE INVENTION
[0003] As used herein, the term "supersorbent" shall mean a sorbent as taught in United States Patent No. 5,779,464 entitled "Calcium Carbonate Sorbent and Methods of Making and Using Same", the teachings of which are hereby incorporated by reference.
[0004] As used herein, the term "microporous" shall mean a pore size distribution of less than 5 nanometers. As used herein, the term "mesoporous" shall mean a pore size distribution of from about 5 nanometers to about 20 nanometers. [0005] Atmospheric CO2 concentration has been increasing steadily since the industrial revolution. It has been widely accepted that the while the CO2 concentration was about 280 ppm before the industrial revolution, it has increased from 315 ppmv in 1959 to 370 ppmv in 2001 [Keeling, CD. and TP. Whorf. 2002. Atmospheric CO2 records from sites in the SIO air sampling network. In Trends: A Compendium of Data on Global Change. Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory, U.S. Department of Energy, Oak Ridge, Tenn., U.S.A. This data is also available from http://cdiac.esd.ornl.gov/ftp/maunaloa- co2/maunaloa.co21. Rising CO2 concentrations has been reported to account for half of the greenhouse effect that causes global warming [IPCC Working Group I. IPCC Climate Change 1995 - The Science of Climate Change: The Second Assessment Report of the Intergovernmental Panel on Climate Change; Houghton, J. T., Meira Filho, L.G., Callander, B.A., Harris, N., Kattenberg, A., Maskell K, Eds.; Cambridge University Press: Cambridge, U.K., 1996]. Although the anthropogenic CO2 emissions are small compared to the amount of CO2 exchanged in the natural cycles, the discrepancy between the long life of CO2 in the atmosphere (50-200 years) and the slow rate of natural CO2 sequestration processes leads to CO2 build up in the atmosphere. The IPCC (Intergovernmental Panel on Climate Change) opines that "the balance of evidence suggests a discernible human influence on the global climate." Therefore, it is necessary to develop cost effective CO2 management schemes to curb its emission. [0006] Many of the envisaged CO2 management schemes consist of three parts - separation, transportation and sequestration of CO2 [FETC Carbon
Sequestration R&D Program Plan: FY 1999-2000. National Energy Technology Laboratory, Department of Energy, Washington, DC, 1999]. The cost of separation and compression of CO2 to 110 bar (for transportation of CO2 in liquid state) is estimated at $30-50 per ton CO2, and transportation and sequestration would cost about $1-3 per ton per 100 km and $1-3 per ton of CO2, respectively [Wallace, D. Capture and Storage of CO2. What Needs To Be Done. Presented at the 6th Conference of the Parties, COP 6, to the United Nations Framework Convention on Climate Change; The Hague, The Netherlands, Nov 13-24, 2000; www.iea.org/envissu/index.htm1. The capture of CO2 imposes severe energy penalties thereby reducing the net electricity output by as much as 13-37% [Herzog, H.; Drake, E.; Adams, E. CO2 Capture, Reuse, and Storage Technologies for Mitigating Global Climate Change. A White Paper; Final Report No. DE-AF22- 96PC01257, Jan 1997]. The dominating costs associated with the current CO2 separation technologies necessitate development of economical alternatives. [0007] Historically, CO2 separation was motivated by enhanced oil recovery
[Kaplan, L. J. Cost-Saving Processes Recovers CO2 from Power-Plant Flue gas. Chem.Eng. 1982, 89 (24), 30-31 ; Pauley, C. P.; Smiskey, P. L.; Haigh, S. N-ReN Recovers CO2 from Flue Gas Economically. Oil Gas J. 1984, 82(20), 87-92]. Currently, industrial processes such as limestone calcination, synthesis of ammonia and hydrogen production require CO2 separation. Absorption processes employ physical and chemical solvents such as Selexol and Rectisol, MEA and KS-2 [Reimer, P.; Audus, H.; Smith, A. Carbon Dioxide Capture from Power Stations. IEA Greenhouse R&D Programme, www.ieagreen.org.uk, 2001. ISBN 1 898373 15 9; Blauwhoff, P.M. M.; Versteeg, G. F.; van Swaaij, W. P. M. A study on the reaction between CO2 and alkanoamines in aqueous solution. Chβm. Eng. Sc/.1984, 39(2), 207-225. Mimura, T.; Simayoshi, H.; Suda, T.; lijima, M.; Mitsuake, S. Development of Energy Saving Technology for Flue Gas Carbon Dioxide Recovery by Chemical Absorption Method and Steam System in Power Plant. Energy Convers. Mgmt. 1997, 38, Suppl. P.S57-S62]. Adsorption systems capture CO2 on a bed of adsorbent materials such as molecular sieves and activated carbon [Kikkinides, E.S.; Yang, R.T.; Cho, S. H. Concentration and Recovery of CO2 from flue gas by pressure swing adsorption. Ind. Eng. Chem. Res. 1993, 32, 2714-2720]. CO2 can also be separated from the other gases by condensing it out at cryogenic temperatures. Polymers, metals such as palladium, and molecular sieves are being evaluated for membrane based separation processes [Reimer, P.; Audus, H.; Smith, A. Carbon Dioxide Capture from Power Stations. IEA Greenhouse R&D Programme, www.ieagreen.org.uk, 2001. ISBN 1 898373 15 9].
[0008] Reaction based processes, as promulgated in this work, can be applied to separate CO2 from gas mixtures. This process is based on a heterogeneous gas- solid non-catalytic carbonation reaction where gaseous CO2 reacts with solid metal oxide (represented by MO) to yield the metal carbonate (MCO3). The reaction can be represented by:
MO + CO2 ^ MCO3 (1)
Once the metal oxide has reached its ultimate conversion, it can be thermally regenerated to the metal oxide and CO2 by the calcination of the metal carbonate product. The calcination reaction can be represented by:
MCO3 > MO + CO2 (2)
As an example of the above-mentioned scheme, Figure 1 shows the variation in the free energy of the carbonation reaction as a function of temperature for calcium oxide. From the figure, we can see that the carbonation reaction is thermodynamically favored with a decrease in temperature (Gibbs free energy declines with a decrease in temperature). However, at lower temperatures, the carbonation reaction is kinetically slow. In fact, it takes geological time scales for the formation of CaCO3 by the reaction between CaO and atmospheric CO2 (at 280-360 ppm) at ambient temperatures. It should also be noted that the carbonation reaction would be favored as long as the free energy is negative. This creates an upper bound of 890 0C for carbonation to occur under a CO2 partial pressure of 1 atm. The equilibrium temperature for this reaction is a function of the partial pressure Of CO2. A reaction based CO2 separation process offers many advantages. Under ideal conditions, MEA captures 6Og CO2/kg, silica gel adsorbs 13.2g CO2/kg and activated carbon adsorbs 88g CO2/kg. The sorption capacity of some metal oxides (such as the modified CaO, presented in this study) is about 70Og CO2/kg of CaO. This is about an order of magnitude higher than the capacity of adsorbents/solvents used in other CO2 separation processes and would significantly reduce the size of the reactors and the material handling associated with CO2 separation.
[0009] Numerous metal oxides exhibit the carbonation and calcination reaction. The calcination temperature of a few metal carbonates (CaCO3 -750 0C, MgCO3 -385 0C, ZnCO3 -340 0C, PbCO3 -350 0C, CuCO3 -225-290 0C and MnCO3 -440 0C) makes them viable candidates for this process. Apart from CaO, gas-solid carbonation of other metal oxides has not been widely studied. The carbonation of ZnO to ZnCO3 at 8-13 0C was low when exposed to CO2 and H2O for over 100 days (Sawada, Y.; Murakami, M.; Nishide, T. Thermal analysis of basic zinc carbonate. Part 1. Carbonation process of zinc oxide powders at 8 and 13 0C. Thermochim. Acta. 1996, 273, 95-102.). MnCO3 undergoes a more complex thermal degradation phenomena. MnCO3 first decomposes to MnO2 at 300 0C, which in turn changes to Mn2O3 at 440 0C. At higher temperatures (-900 0C), the final thermal decomposition product was identified as Mn3O4 (Shaheen, W. M.; Selim, M. M. Effect of thermal treatment on physicochemical properties of pure and mixed manganese carbonate and basic copper carbonate. Thermochim. Acta. 1998, 322(2), 117-128.). Different oxides of manganese provide the flexibility of exploiting the carbonation/calcination reaction over a wider temperature range. Aqueous phase MgO carbonation has been studied for its suitability for mineral-based CO2 sequestration (Fernandez, A.I.; Chimenos, J. M.; Segarra, M.; Fernandez, M.A.; Espiell, F. Kinetic study of carbonation of MgO slurries. Hydrometallurgy. 1999, 53, 155-167). The carbonation extent of Mg(OH)2 was about 10% between 387-400 0C and 6% formation between 475-500 0C (Butt, DP.; Lackner, K.S.; Wendt, C.H.; Conzone, S.D.; Kung, H.; Lu, Y- C; Bremser, J. K. Kinetics of Thermal Dehydroxylation and Carbonation of Magnesium Hydroxide. J. Am. Ceram. Soc.1996, 79(7), 1892-1898). They attributed the low conversions to the formation of a non-porous carbonate product layer. This layer hinders the inward diffusion of CO2 and the outward diffusion of H2O (a product of the carbonation reaction) leading to low conversions. The carbonation of PbO was studied as a part of the chemical heat pump process (Kato, Y.; Saku, D.; Harada, N.; Yoshizawa, Y. Utilization of High Temperature Heat from Nuclear Reactor using Inorganic Chemical Heat Pump. Progress in Nuclear Energy. 1998, 32(3-4), 563- 570. & Kato, Y.; Harada, N.; Yoshizawa, Y. Kinetic feasibility of a chemical heat pump for heat utilization from high temperature processes. Applied Thermal Engineering. 1999, 19, 239-254). They reported 30% conversion in an hour under 100% CO2 atmosphere at 3000C. Furthermore, they found the reactivity of PbO to drop with the number of carbonation-calcination cycles. [0010] Carbonation of calcium oxide has been widely studied. Related applications of the CaO carbonation and calcination include the storage of energy (Barker, R. The Reversibility of the Reaction CaCO3 = CaO + CO2. J. Appl. Chem. Biotechnol. 1973, 23, 733-742) and the zero emission coal alliance process, consisting of hydrogasification of coal fueled by the heat of the carbonation reaction (Tinkler, M.J.; Cheh, C. Towards a Coal-capable Solid Oxide Fuel Cell System. Proceedings of the 26th International Technical Conference on Coal Utilization and Fuel Systems; Clearwater, Florida, March 5-8, 2001 ; pp 569-570). The gas-solid CaO-CO2 reaction proceeds through two rate-controlling regimes. The first regime involves a rapid, heterogeneous chemical reaction. In the second regime, the reaction slows down due to the formation of an impervious layer of CaCO3. This product layer prevents the exposure of unreacted CaO in the particle core to CO2 for further carbonation. The kinetics of the second regime is governed by the diffusion of ions through the CaCO3 product layer. The activation energy was estimated to be 21 kcal/mol below 688 K and 43 kcal/mol above it for the product layer diffusion, based on the counter migration of CO3 2" and O2" ions through the product layer (Bhatia, S. K.; and Perlmutter, D. D. Effect of the product layer on the kinetics of the CO2-Lime Reaction. AIChE J. 1983, 29(1), 79-86).
[0011] The extent of the carbonation reaction reported in many studies has also shown considerable variation. Stoichiometrically, 56g of CaO should react with 44g of CO2 to form 10Og of CaCO3. This translates to about 78.6-wt% capacity for CaO. However, the structural limitations prevent the attainment of theoretical conversion. The extent of carbonation was only 23-wt% in 30 minutes at 600 0C (Dedman, AJ. ; Owen, A.J. Calcium Cyanamide Synthesis, Part 4.- The reaction CaO + CO2 = CaCO3. Trans. Faraday Soc.1962, 58, 2027-2035). A higher surface area CaO sorbent provided 55-wt% CO2 sorption (Bhatia, S. K.; and Perlmutter, D. D. Effect of the product layer on the kinetics of the CO2-l_ime Reaction. AIChEJ. 1983, 29(1), 79-86). 64-wt% CO2 sorption was achieved at 1050 0C temperature and 11.74 atm CO2 pressure in 32 hours (Mess, D.; Sarofim, A.F.; Longwell, J. P. Product Layer Diffusion during the Reaction of Calcium Oxide with Carbon Dioxide. Energy and Fuels. 1999, 13, 999-1005). However, the extent of carbonation at lower temperature/pressure conditions that are more characteristic of CO2 containing gaseous mixtures is absent in their work. The limitation in total conversion stems essentially from the nature of the initial pore size distribution of the CaO sorbent. Microporous sorbents (pore size < 2 nm) are very susceptible to pore blockage and plugging due to the formation of higher molar volume product (molar volume of CaO: 17 cm3/mol; molar volume of CaCOa: 37 cnrϊVmol). CaO sorbents obtained from naturally occurring precursors are usually microporous in nature. At the end of the kinetically controlled regime, diffusion processes through the product layer control the reaction rate. Similar structural limitations have prevented calcium-based sorbents from attaining theoretical conversion for the sulfation reaction between CaO and sulfur dioxide (SO2) as well (Wei, S.-H.; Mahuli, S.K.; Agnihotri, R.; Fan, L.-S. High Surface Area Calcium Carbonate: Pore Structural Properties and Sulfation Characteristics. Ind. Eng. Chem. Res. 1997, 36(6), 2141-2148). They suggested that a mesoporous structure, which maximizes porosity in the 5-20 nm pore size range, would be less susceptible to pore pluggage. This structure would also be able to provide sufficient surface area to ensure rapid kinetics. Their modified precipitation technique resulted in a mesoporous CaCO3 structure that also had a high BET surface area determined by nitrogen (60 m2/g). A similar approach could also enhance the reactivity of CaO sorbents towards the carbonation reaction, which is the focus of this study.
[0012] Lastly, it is important that the CaO sorbents maintain their reactivity over many carbonation and calcination cycles. The conversion of CaO dropped from about 73% in the first carbonation cycle to 43% at the end of the 5th cycle at 866 0C (Barker, R. The Reversibility of the Reaction CaCO3 = CaO + CO2. J. Appl. Chem. Biotechnol. 1973, 23, 733-742 & Barker, R. The Reactivity of Calcium Oxide Towards Carbon Dioxide and its use for Energy Storage. J. Appl. Chem. Biotechnol. 1974, 24, 221-227). Barker suggested that the CaCO3 layer is about 22 nm thick and his latter work showed repeated 93% conversion over 30 cycles at 629 0C on 10 nm CaO particles. In another study, cyclical studies conducted at a carbonation temperature of 880 0C and calcination at 860 0C led to a drop in conversion from 70% in the first carbonation to 38% in the 7th carbonation step (Kato, Y.; Harada, N.; Yoshizawa, Y. Kinetic feasibility of a chemical heat pump for heat utilization from high temperature processes. Applied Thermal Engineering. 1999, 19, 239-254). The process described here leads to > 95% conversion due to the application of novel mesoporous CaO sorbents for CO2 capture and maintains their reactivity over repeated cycles of carbonation and calcination.
Enhanced Hydrogen Production Integrated with CO2 Separation in a Single-Stage Reactor
[0013] There has been a global push towards the development of a hydrogen economy. The main premise behind this drastic alteration in our energy usage stems from the fact that the use of hydrogen in portable and mobile applications would be the most environmentally beneficial process that leads only to the emission of water. However, the biggest issue that needs to be addressed for the success of the hydrogen-based economy involves the source of hydrogen itself. While hydrogen may be considered as the best "carrier" of energy, there is clearly no hydrogen "wells" on earth. The major processes for hydrogen production from fossil fuels consist of steam reforming of methane (SMR), coal gasification, catalytic cracking of natural gas, and partial oxidation of heavy oils. Other processes consist of water electrolysis, thermo chemical water decomposition, biological processes, etc. (Rosen and Scott, 1998; Rosen, 1996). However, water electrolysis is not a very energy efficient process. [0014] Water gas, a mixture of CO, CO2, H2O and H2, is formed by the gasification of coal by sub-stoichiometric air and/or steam. Irrespective of the initial concentration of these four gases, the reversible water gas shift (WGS) reaction gets initiated until the exact ratio of the concentration of these gases reaches a particular equilibrium constant KWGS that is a function of temperature. The WGS reaction and its equilibrium constant can be written as:
WGS Reaction: CO + H2O <=> CO2 + H2 ΔH = - 40.6 kJ/mol (1 )
WGS equilibrium constant:
Figure imgf000012_0001
where T is in 0C. From equation (2), it can be observed that KWGS reduces with increasing temperature. This means that processes aimed at converting coal-derived gas to hydrogen at high temperatures are thermodynamically restricted. While catalysts aid in achieving this equilibrium, they cannot alter the value of K to provide a higher hydrogen yield. An effective technique to shift the reaction to the right for enhanced hydrogen generation has been to remove hydrogen from the reaction mixture. This premise has lead to the development of hydrogen separation membranes. However, membranes cannot completely remove hydrogen from the mixture. Any remaining hydrogen would dilute CO2 after its utilization in either a fuel cell or gas turbine. [0015] Another option for driving the WGS reaction forward is to remove CO2 from the reaction mixture by reacting it with CaO. The carbonation reaction can be written as:
Carbonation Reaction: CaO + CO2 -> CaCO3 (ΔH = - 183 kJ/mol) (3) Under the appropriate reaction temperature, CO2 concentration can be lowered down to ppm levels by reaction (3), thereby enabling the maximum production of hydrogen from carbon via reaction (1). By conducting the reaction such that CO is the limiting reactant, we can ensure complete utilization of the fuel as well. Besides these advantages, CO2 is simultaneously removed from the gas mixture in the form of CaCO3, thereby improving the purity of the hydrogen stream (the other contaminant being only water). The spent sorbent can then be calcined separately to yield pure CO2 stream, which is then amenable for compression and liquefaction before its transportation to sequestration sites. Calcination reaction, reverse of the carbonation reaction can be written as: Calcination Reaction: CaCO3 → CaO + CO2 (ΔH = + 183 kJ/mol) (4) The resulting CaO sorbent is recycled to capture CO2 in the next cycle. This cyclical CCR process can be continued so long as the sorbent provides a satisfactory CO2 capture.
[0016] To obtain high purity H2, the WGS reaction is generally carried out in
0 two stages for: (1) high temperature shift (250-500 C) using iron catalysts and (2)
O low temperature shift (210-270 C) using copper-based catalysts (Gerhartz, 1993; Bohlbro, 1969). Copper based catalysts are extremely intolerant to small quantities of sulfur (< 0.1 ppm) and hence the fuel gases need to be desulfurized upstream of the WGS reactor. Besides, to achieve satisfactory carbon monoxide conversion a considerable quantity of high-pressure steam is required. For example, to lower the CO content of the typical fuel gas from 45 % (inlet) to 3% (outlet) a total steam
addition of 1.18 kg/m of the gas is required, at a total pressure of 60 bar and 410 C
O
(Gerhartz, 1993). The steam to CO ratio at 550 C can be as high as 50 during a single-stage operation or 7.5 for a more expensive dual-stage process to obtain 99.5 % pure H2 (David, 1980). This is necessary due to the equilibrium limitation inherent in the WGS reaction. From the point of view of H2 production, even though higher temperatures lead to improved kinetics, WGS has poor equilibrium conditions at the higher temperatures. However, the continuous removal of the carbon dioxide product from the reaction chamber will incessantly drive the equilibrium-limited water-gas shift reaction forward. This will ensure a high yield and purity of hydrogen with near stoichiometric amounts of steam needed for the reaction. Besides, the reaction can now be carried out at higher temperatures leading to superior kinetics in the forward direction. Thus the major equilibrium related drawback in this process could be overcome. The continuous CO2 removal can be brought about by the carbonation reaction of a metal oxide to give the corresponding metal carbonate. We have identified a high reactivity, mesoporous calcium oxide as the potential sorbent for the in-situ CO2 capture given by eqn. 3.
[0017] The success of this process would effectively bridge coal gasification to fuel cell usage and chemical synthesis. Other side benefits of this process involve the potential for removal of sulfur and heavy metals such as arsenic and selenium from the fuel gas stream.
[0018] Recently, Harrison and co-workers reported a single-step sorption- enhanced process to produce hydrogen from methane (Balasubramanian et al., 1999; Lopez Ortiz and Harrison, 2001). They used the traditional concept of SMR with WGS using Ni-based catalyst to produce hydrogen, coupled with this novel scheme of in-situ continuous CO2 capture using a calcium-based dolomite sorbent. They obtained high hydrogen yields with 97 % purity (dry basis). [0019] However, they reported a low "calcium" conversion in the sorbent of about 50% at the beginning of the breakthrough to about 83% at the end of the test. These conversion calculations are based on only the calcium portion of their dolomite sorbent. Their total sorbent conversion will be much lower than these values as dolomite does not entirely contain calcium based material. In fact, dolomite comprises of nearly 50wt.% calcium, which participates in the reaction to some extent, and the remaining portion of the sorbent (mainly magnesium oxide) stays unreacted. Further, they attribute the incomplete conversions of the calcium material to the concept of pore filling and pluggage at the pore-mouths of these sorbent particles by CaCO3 product layer, preventing the access of CO2 in the gas to unreacted CaO surface at the pore interiors.
[0020] Harrison and co-workers regenerated the dolomite sorbent in streams of N2, 4%O2 in N2 and pure CO2. They had to use high regeneration temperatures of
O
800-950 C, especially while using pure CO2. Exposure of the reforming catalyst to an oxidizing atmosphere {viz. O2/N2 or CO2) while regenerating the sorbent used to oxidize the Ni catalysts to NiO. Hence, the catalyst had to be reduced back to Ni before every cycle or the sorbent-catalyst mixture had to be separated after every run so that only the sorbent is subjected to the regeneration conditions. Further, the temperature of operation can be lowered by regeneration in a pure N2 stream. However, it would not solve the problem of CO2 separation due to the formation of a CO2/N2 gas mixture. Calcination in a pure CO2 stream will result in higher operating temperatures due to the thermodynamic limitations of the calcination reaction in presence of the CO2 product. Higher temperatures and the presence of CO2 during calcination would cause the sorbent to sinter. This is in agreement with the results of multiple carbonation-calcination cycle tests for dolomite by Harrison and co-workers (Lopez Ortiz and Harrison, 2001) in pure CO2 stream (800-950 0C). They observed a decrease in "calcium" conversion from 83 % in the 1st cycle to about 69 % in the 10th cycle itself. However, a mesoporous high suface area calcium based sorbent (precipitated calcium carbonate, PCC) developed at OSU has undergone 100 cycle experiments. The PCC sorbent has shown 85% conversion in the 1st cycle 66.7% in the 10th cycle and 45.5% in the 100th cycle towards carbonation. These experiments were carried out in a TGA at 7000C in a 10% CO2 stream in the carbonation cycle and 100% N2 gas in the calcination cycle, with 30 minute residence times for each cycle. Therefore this project aims testing this PCC based sorbent towards further enhancing the WGSR and overcoming some of the problems faced by Harrison and co-workers.
SUMMARY OF THE INVENTION [0021] The present invention includes a calcium oxide, its usage for the separation of CO2 from multicomponent gas mixtures and the optimum process conditions necessary for enhancing the repeatability of the process. [0022] A preferred method for separating carbon dioxide from a flow of gas comprising carbon dioxide comprises the steps of: (1) directing the flow of gas to a gas-solid contact reactor, the gas-solid contact reactor contains at least one sorbent comprising at least one metal oxide; (2) reacting the carbon dioxide with the at least one sorbent so as to remove the carbon dioxide from said flow of gas, thereby converting the at least one sorbent into spent sorbent; (3) calcining the spent sorbent so as to liberate the carbon dioxide from the spent sorbent, thereby regenerating the sorbent; and (4) repeating the aforementioned steps.
[0023] Although any metal oxide may be employed, it is preferred that the at least one metal oxide is selected from the group consisting of: ZnO, MgO, MnO2, NiO, CuO, PbO, and CaO. Further, it is preferred that the spent sorbent is a metal carbonate. [0024] It is preferred that the sorbent has a sorption capacity of at least about
70 grams of carbon dioxide per kilogram of sorbent. However, it is even more preferred that the sorbent has a sorption capacity of at least about 300 grams of carbon dioxide per kilogram of sorbent. Irrespective of the sorption capacity of the sorbent, it is preferred that the sorbent has substantially the same sorption capacity after calcining as the sorbent had prior to adsorbing the carbon dioxide. [0025] Although any calcination method may be employed, it is preferred that the calcining is performed under at least partial vacuum. It is also preferred that the calcining is performed by steam. [0026] The present invention includes facilities practicing the aforementioned method.
[0027] A method for separating carbon dioxide from a flow of gas comprising carbon dioxide of the present invention comprises the steps of: (1 ) directing the flow of gas to a first gas-solid contact reactor, the first gas-solid contact reactor containing at least one sorbent, the sorbent comprising at least one metal oxide; (2) reacting the carbon dioxide in the flow of gas on the sorbent in the first gas-solid contact reactor so as to remove the carbon dioxide from the flow of gas; (3) directing the flow of gas to a second gas-solid contact reactor when the sorbent in the first gas-solid contact reactor is spent thereby forming spent sorbent, the second gas-solid contact reactor containing at least one sorbent, the sorbent comprising at least one metal oxide; (4) reacting the carbon dioxide in the flow of gas on the sorbent in the second gas-solid contact reactor so as to remove the carbon dioxide from the flow of gas; (5) calcining the spent sorbent from the first gas-solid contact reactor so as to generate carbon dioxide and to regenerate the sorbent; (6) directing the flow of gas to the first gas- solid contact reactor when the sorbent in the second gas-solid contact reactor is spent, thereby forming spent sorbent; and (7)calcining the spent sorbent from the second gas-solid contact reactor so as to generate carbon dioxide and to regenerate the sorbent.
[0028] Although any calcination method may be employed, it is preferred that the calcining is performed under at least partial vacuum. It is also preferred that the calcining is performed by steam. This applies to both gas-solid contact reactors.
[0029] Although any metal oxide may be utilized, it is preferred that the at least one metal oxide is selected from the group consisting of: ZnO, MgO, MnO2,
NiO, CuO, PbO, and CaO. [0030] It is preferred that the sorbent has a sorption capacity of at least about
70 grams of carbon dioxide per kilogram of sorbent. However, it is even more preferred that the sorbent has a sorption capacity of at least about 300 grams of carbon dioxide per kilogram of sorbent. Irrespective of the sorption capacity of the sorbent, it is preferred that the sorbent has substantially the same sorption capacity after calcining as the sorbent had prior to adsorbing the carbon dioxide.
[0031] The present invention also includes facilities practicing the aforementioned method
[0032] A method for regenerating a spent sorbent for carbon dioxide of the present invention comprises the steps of: (1) providing a spent sorbent, the spent sorbent comprising metal carbonate; and (2) calcining the spent sorbent so as to liberate carbon dioxide gas and so as to regenerate the spent sorbent thereby forming a sorbent comprising a metal oxide.
[0033] It is preferred that the spent sorbent is calcium carbonate. It is further preferred that the metal oxide is calcium oxide. [0034] It is preferred that the sorbent has substantially the same sorption capacity after calcining as the sorbent had prior to adsorbing the carbon dioxide. [0035] Although any calcination method may be employed, it is preferred that the calcining is performed under at least partial vacuum. It is also preferred that the calcining is performed by steam. This applies to both gas-solid contact reactors. [0036] The present invention includes facilities practicing the aforementioned method.
[0037] A method for producing a sorbent of the present invention comprises the steps of: (1 ) obtaining a structurally altered high surface area calcium carbonate having a surface area of at least 25.0 m2/g, a pore volume of at least 0.05 cm3/g, and a mesoporous pore size distribution; and (2) calcining the structurally altered high surface area calcium carbonate so as to produce a sorbent having a surface area of less than 22 m2/g, a pore volume of at least 0.005 cm3/g, and a mesoporous pore size distribution. [0038] Although any calcination method may be employed, it is preferred that the calcining is performed under at least partial vacuum. It is also preferred that the calcining is performed by steam. This applies to both gas-solid contact reactors. [0039] The present invention includes sorbents made according to the aforementioned method. [0040] A sorbent according to the present invention comprising calcium oxide having a surface area of at least 12.0 m2/g and a pore volume of at least 0.015 cm3/g, the calcium carbonate sorbent having sorption capacity of at least about 70 grams of carbon dioxide per kilogram of sorbent. [0041] In addition to the novel features and advantages mentioned above, other objects and advantages of the present invention will be readily apparent from the following descriptions of the drawing(s) and preferred embodiment(s).
BRIEF DESCRIPTION OF THE DRAWINGS [0042] Figure 1 depicts the Gibbs Free Energy diagram for the carbonation reaction, CaCθ3→CaO + CO2, as a function of temperature.
[0043] Figure 2 illustrates the performance of calcium oxide for the carbonation reaction.
[0044] Figure 3 compares the XRD diffractograms of CaO derived from various precursors.
[0045] Figure 4 is a schematic diagram of a carbonator reactor for the synthesis of precipitated calcium carbonate.
[0046] Figure 5 shows the change in the pH of the slurry as a function of
Ca(OH)2 loading. (500 ml_ water, 0.0575% N40V® dispersant, 4 scfh CO2). [0047] Figure 6 depicts the effect of Ca(OH)2 loading on the morphology of
Precipitated Calcium Carbonate (PCC) (500 mL water, 0.0575% N40V® dispersant, 4 scfh CO2).
[0048] Figure 7 compares the pore size distribution of four CaO precursors.
[0049] Figure 8 compares the conversion of four CaO sorbents under pure CO2 at 6500C.
[0050] Figure 9 illustrates the effect of temperature on the carbonation of
PCC-CaO.
[0051] Figure 10 illustrates the carbonation-calcination cycles on Aldrich
CaCO3 and PCC at 700 0C. [0052] Figure 11 shows extended carbonation-calcination cycles on precipitated calcium carbonate (PCC) powder at 700 0C.
[0053] Figure 12 compares the effect of initial surface area of PCC-CaO to its reactivity towards the carbonation reaction at 700 0C. [0054] Figure 13 depicts the effect of vacuum calcination on the reactivity of
PCC-CaO towards the carbonation reaction at 700 0C.
[0055] Figure 14 provides a flow sheet depicting the integration of the current process in the overall coal-gasifier electric production facility.
[0056] Figure 15 illustrates thermodynamic data for predicting the temperature zones for hydration and carbonation of CaO.
[0057] Figure 16 illustrates thermodynamic data for predicting the equilibrium
H2S concentration for CaO sulfidation with varying steam concentration (PTotal - 1 atm).
[0058] Figure 17 shows a modified reactor set-up with steam generating unit for investigating WGS and carbonation reactions.
[0059] Figure 18 illustrates the set-up for combined vacuum/sweep gas calcination experiments allowing the use of larger sorbent samples.
[0060] Figure 19 is a pore size distribution of the HTS and LTS obtained from
BET analysis. [0061] Figure 20 shows the pore size distribution of various calcium oxide precursors.
[0062] Figure 21 shows the effect of reaction temperature on the CO conversion (0.5 g HTS catalyst, 3% CO, H2O/CO ration = 3, total flow = 1.5 slpm).
[0063] Figure 22 shows the extent of reaction equilibrium as a function of temperature for the WGS reaction. [0064] Figure 23 is a breakthrough curve of CO conversion using a PCC-HTS catalyst system (T=600C, 3% CO, 9% H2O, Total flow = 1.5 slpm).
[0065] Figure 24 is a breakthrough curve of CO conversion using a LH-HTS catalyst system (T=600C, 3% CO, 9% H2O, total flow = 1.5 slpm). [0066] Figure 25 provides a comparison of breakthrough curves for PCC-HTS and LH-HTS systems (T=600C, 3% CO, 9% H2O, Total flow = 1.5 slpm).
[0067] Figure 26 depicts a typical steam generation scenario and use.
[0068] Figure 27 depicts one implementation of one embodiment of the present invention. [0069] Figure 28 depicts one implementation of one embodiment of the present invention.
[0070] Figure 29 depicts one implementation of one embodiment of the present invention.
[0071] Figure 30 depicts one implementation of one embodiment of the present invention.
[0072] Figure 31 depicts one implementation of one embodiment of the present invention.
[0073] Figure 32 depicts one implementation of one embodiment of the present invention. [0074] Figure 33 depicts one implementation of one embodiment of the present invention.
[0075] Figure 34 depicts one implementation of one embodiment of the present invention.
[0076] Figure 35 depicts one implementation of one embodiment of the present invention. [0077] Figure 36 depicts one implementation of one embodiment of the present invention.
[0078] Figure 37 depicts one implementation of one embodiment of the present invention. [0079] Figure 38 illustrates the calcination and subsequent sulfation of the precipitated calcium carbonate.
[0080] Figure 39 provides an XRD plot obtained after drying the product.
[0081] Figure 40 provides an XRD plot of the solid product after 10 minutes of carbonation. [0082] Figure 41 provides an XRD plot of the solid product after 60 minutes of carbonation.
[0083] Figure 42 compares the pore size distribution for two CaCO3 products.
DETAILED DESCRIPTION OF THE PREFERRED EMBODIMENT(S)
[0084] In accordance with the foregoing summary, the following presents a detailed description of the preferred embodiment(s) of the invention that are currently considered to be the best mode.
Chemicals, Sorbents and Gases
[0085] Naturally occurring limestone (CaCOa) and hydrated lime (Ca(OH)2), synthesized from it were obtained from Linwood Mining and Minerals. Dolomite
(CaCO3*MgCO3) was procured from the National Dolomite Company. The purity of
these ores was above 90%. High purity metal oxides such as ZnO1 MgO, MnO2, NiO,
CuO, PbO, CaO were obtained from Aldrich Chemical Company. Precipitated calcium carbonate (PCC) was synthesized from Linwood hydrate by the procedure described in a following section. N40V® dispersant, a sodium salt of a carboxylic acid, used in the synthesis of PCC was obtained from Allied Colloid. The synthesis procedure is described in detail in a following section. N2 and CO2 used for calcination and carbonation experiments were 99.999% and 99.9% pure, respectively.
Sorbent Reactivity Testing and Structural Analysis [0086] The reactivity testing of CaO sorbents for carbonation was carried out in a Perkin Elmer Thermogravimetric Analyzer (TGA-7) apparatus. The balance can accurately measure up to 1 microgram. A small sample of the sorbent (5-20 mg) is placed in a quartz boat. The weight of the sample was recorded every second. The structural properties of CaO sorbents and their precursors were tested in a NOVA 2200 analyzer (Quantachrome Company). The BET surface area, pore volume, and pore size distribution were measured at -1960C using nitrogen as the adsorbent.
Screening of Metal Oxides
[0087] Metal oxides such as ZnO, MgO, CuO, MnO2, NiO, PbO and CaO that undergo the CCR scheme in the 800-200 0C temperature range were analyzed for their reactivity in a TGA. A powdered sample of these oxides was placed in a quartz pan and pure CO2 was passed over the sample metal oxide. The temperature was then slowly raised and the weight of the sample was continuously monitored. An increase in the weight of the sample is an indication of the formation of metal carbonate. Figure 2 provides experimental data for the carbonation of lime (Ca(OH)2) under flowing pure CO2 gas. With an increase in temperature, the weight of the sample increases till the temperature reaches about 8900C. Calcination, which is thermodynamically favored above 890 0C at 1 atm CO2 partial pressure, causes a rapid decrease in weight until the sorbent converts completely to CaO. When the sample is reheated, the weight starts to increase again and the process is repeated once more. Besides proving that CaO is a viable candidate, the data also shows recyclability of the sorbent.
XRD Analysis of CaO obtained from its precursors: [0088] CaO was identified as a viable candidate for the carbonation- calcination reactions. However, a variety of precursors can be calcined to obtain the CaO sorbents necessary for the carbonation reaction. Common and economical precursors include calcium carbonate, calcium hydroxide and dolomite. The other important source of CaO is via the calcination of synthesized high surface area precipitated calcium carbonate. In order to compare the crystal structure of the CaO sorbents obtained from these sources, XRD patterns were obtained on all the CaO sorbents. Figure 3 depicts these d iff ractog rams (a. Calcined Aldrich-CaO; b. Dolomite-CaO; c. Ca(OH)2-CaO); d. PCC-CaO; e. Limestone-CaO; and f. Aldrich- CaO). From this figure we can conclude that the crystal structure of the CaO sorbents obtained from numerous sources is identical. Only the XRD pattern corresponding to dolomite-derived CaO shows extra peaks due to the presence of MgO in the calcined dolomite. Based on the similarity in all the CaO structures, it can be assumed that any difference in reactivity of CaO for carbonation is an artifact of the sorbent morphology and not due to the chemistry of the gas-solid reaction that occurs on the CaO surface. Precipitated Calcium Carbonate (PCC) synthesis
[0089] Structurally altered high surface area CaO precursors were synthesized based on the procedure outlined elsewhere (Fan, L-S.; Ghosh-Dastidar, A.; Mahuli, S.; Calcium Carbonate Sorbent and Methods of Making the Same. US Patent # 5,779,464 and Agnihotri, R.; Chauk, S.; Mahuli, S.; Fan, L.-S. Influence of Surface Modifiers on Structure of Precipitated Calcium Carbonate. Ind. Eng. Chem. Res. 1999, 38, 2283-2291 ). A schematic diagram of the slurry bubble column used for this purpose is shown in Figure 4. The carbonator 40 consists of a 2" OD Pyrex tube 40a. A porous frit 4Od at the bottom, disposed over glass beads 4Of, provides good distribution of CO24Og through the slurry 40c. A K-type thermocouple 4Oh inserted in the slurry continuously records the slurry temperature. A pH probe 40b monitors the pH of the slurry as the reaction medium changes from a basic to an acidic solution as the reaction proceeds. First, 500 ml of distilled water is poured into the carbonator, followed by the addition of 0.0575g of N40V®. 12.8g of Ca(OH)2 is added to the solution to provide a loading of 2.56% by weight. This corresponds to a concentration of 16-sat (concentration of Ca(OH)2 is 16 times its saturation solubility limit). The solubility of Ca(OH)2 (- 0.16 g/100g water) leads to a pH of 12 at the start of the experiment. The remaining Ca(OH)2 remains suspended in the solution. The ratio of N40V® and Ca(OH)2 loading is chosen to create a surface charge of zero on the incipiently formed CaCO3 particles. The flow of CO24Oe into the carbonator is then started and the pH was continuously monitored. Figure 5 shows the change in pH with reaction time as a function of Ca(OH)2 loading. CO2 dissolved in water provides carbonate ions that react with Ca++ ions to form CaCO3 according to the reaction below:
Ca2+ + CO3 2" ■* CaCO3 (3) CaCO3 has a much lower solubility in water (~0.0012 g/10Og water) compared to Ca(OH)2 and thus precipitates out. As the reaction proceeds, Ca2+ ions get depleted, but are continuously replenished by the suspended Ca(OH)2. Hence the pH remains 12. As the reaction proceeds, Ca(OH)2 ultimately gets depleted and the concentration of Ca2+ ions cannot be maintained at its solubility limit. On the other hand, continued dissolution of CO2 gas leads to the accumulation of H+ ions causing the solution to become acidic. Eventually, the pH settles at about 6.0, corresponding to equilibrium solubility of CO2 in water at ambient temperature. This also signals the end of the carbonation of all Ca(OH)2. The slurry is then removed from the precipitator, vacuum filtered and stored in a vacuum oven at 90-110 0C for 20 hours to completely remove the moisture. Higher Ca(OH)2 loading requires more reaction time as evident from Figure 5.
Effect of the ratio of Ca(OH)2 and dispersant on PCC morphology: [0090] Precipitated calcium carbonate can be obtained by the reaction between carbonate and calcium ions in solution. It is known that the CaCO3 nuclei that precipitate out have positive surface charge on them that prevent agglomeration (Agnihotri, R.; Chauk, S.; Mahuli, S.; Fan, L.-S. Influence of Surface Modifiers on Structure of Precipitated Calcium Carbonate. Ind. Eng. Chem. Res. 1999, 38, 2283- 2291). The resulting structure is also microporous in nature. However, the structural properties of the synthesized PCC can be altered by the use of negatively charged dispersants that neutralize the surface charges. This makes the ratio between the Ca(OH)2 loading and the dispersant used very critical. Besides, the effect of Ca(OH)2 loading in the slurry was studied to enhance the productivity of the precipitation process by synthesizing more PCC from the same slurry volume. 8-sat, 16-sat and 24-sat were used as Ca(OH)2 loading levels, all other factors remaining constant. It can be seen from Figure 6 and Table 1 that at a concentration of 8-sat, there is proportionally more dispersant in the slurry causing the incipiently formed CaCO3 particles to be negatively charged. The negative charge prevents the agglomeration of these nuclei eventually leading to the formation of microporous PCC as shown in Figure 6. Its surface area is also relatively lower. At a Ca(OH)2 loading corresponding to 16-sat, the ratio of N40V® and CaCO3 is balanced and the surface charge on the nuclei is zero. This allows optimal association of these nuclei leading to a predominantly mesoporous structure. The SA of PCC under these optimum conditions is also the highest at 38.3 m2/g. As the loading of Ca(OH)2 is raised to 24- sat, there is not enough N40V® dispersant to neutralize the surface charge on all the incipiently formed nuclei. There could possibly be some positively charged particles. This again creates non-optimum conditions leading to a loss in SA and PV compared to the 16-sat case. Another experiment was conducted to process a 32-sat Ca(OH)2 slurry keeping the Ca(OH)2 to N40V® ratio constant. The SA/PV of PCC synthesized from a 32-sat slurry was 37.07 m2/g and 0.139 cm3/g respectively; lending support to the fact that higher mass of PCC can be synthesized from the same amount of slurry.
Figure imgf000028_0001
Table 1 : Morphological properties of PCC as a function of N40V®:Ca(OH)2 loading ratio (500 ml water, 0.0575% N40V® dispersant, 4 scfh CO2).
Pore Structure of CaO sorbents
[0091] CaO sorbents were synthesized by calcining various CaO precursors such as Linwood calcium carbonate (LC), dolomite (DL), Linwood calcium hydroxide (LH), and precipitated calcium carbonate (PCC). For convenience, the oxides derived from these sources are termed as LC-CaO, FCD-CaO (for fully calcined dolomite-CaO), LH-CaO, and PCC-CaO, respectively. The procedure involved heating the precursor in flowing nitrogen beyond the calcination temperature (800- 9500C) for an hour followed by its storage in a desiccator. Structural properties such as surface area (SA) and pore volume (PV) of these chemicals are listed in Table 2 and their pore size distributions are shown in Figure 7. The SA of naturally occurring minerals, LC and dolomite was very low, 1.06 and 1.82 m2/g, respectively. LH was synthesized by first calcining the LC followed by its hydration. LH exhibited a considerably higher SA (13.19 m2/g) and PV compared to the LC. The SA of PCC (38.3 m2/g), however, was the highest among all precursors. From Figure 5, we can infer that the structures of LC, DL and LH are predominantly microporous in nature. Most of the porosity lies in pores below 5 nm in diameter. In contrast, the maximum in PV occurs at 15 nm for PCC and most of its PV originates from mesopores in the 5-25 nm range.
Figure imgf000029_0001
Table 2: Morphological properties (surface area and pore volume) of various CaO sorbents and their precursors.
Carbonation of CaO sorbents
[0092] The performance of these four CaO sorbents was tested in a TGA. The experimental procedure consisted of placing 6-12 mg of the chosen CaO sorbent in a thin layer in a quartz pan to minimize external mass transfer resistances. The sorbent was then heated in flowing nitrogen (5.0 grade, 99.999% pure) to the desired temperature. The representative temperatures used in these experiments were 550 0C, 600 0C and 650 0C. Once the desired temperature was reached, the flow was switched to 100% CO2 stream. The increase in weight with time was recorded and the conversion of CaO to CaCO3 was calculated from the increase in weight. Only the data obtained at 650 0C is reported here. The performance of the four CaO sorbents, LC-CaO, FCD-CaO, LH-CaO and PCC-CaO at 650 0C is depicted in Figure 8. Initially, CO2 diffuses into the pores of the LC-CaO and the reaction takes place on the CaO surface provided by the pores. The figure shows that there is a rapid increase in weight in the first 1-2 minutes. The conversion attained in this kinetically controlled regime depends on the initial surface area of the CaO sorbent. LC-CaO and FCD-CaO attained 40-45% conversion, while LH-CaO and PCC-CaO attained about 60% and 54% conversion, respectively, in this regime. After this regime, conversion increases relatively slowly with time. The increase in conversion is only about 2-4% in the next hour for LC-CaO and FCD-CaO. This confirms the susceptibility of micropores to pore filling and pore pluggage described earlier due to the formation of a higher volume product, CaCO3. The trend is not as dramatic for the case of LH-CaO because of its relatively higher initial surface area. The conversion for LH-CaO increases by another 18% in the diffusion controlled regime. However, the increase in conversion for PCC-CaO is about 34-36% more in the second regime. Since the PCC-CaO structure is mesoporous, the formation of CaCO3 product layer is not able to plug all the pore mouths. This in turn allows the heterogeneous reaction to occur on a larger CaO surface. Once the kinetically controlled regime is over, diffusion of ions occurs through a larger area, ultimately leading to a higher conversion of 88-90% for PCC-CaO. Figure 9 shows the effect of temperature on the carbonation of PCC-CaO. It can be seen that the extent of conversion in the kinetic regime is different at different temperatures. However, unlike LC-CaO, the conversion at any temperature does not seem to taper off and given sufficient time, PCC-CaO is capable of attaining 90% or higher conversion at all of these temperatures.
Cyclic calcination and carbonation [0093] One of the possible hurdles in the utilization of metal oxides for the carbonation and calcination reaction scheme is its vulnerability to sintering due to the thermal cycling imposed by the cyclical nature of these reactions. Cyclical studies were carried out to quantify any loss in reactivity of these sorbents upon multiple cycles. The temperature chosen for cyclical studies was 700 0C. This temperature is sufficient to achieve carbonation in the presence of pure CO2, and also to calcine the CaCO3 so formed after the gas is switched from CO2 to N2. A variety of precursors were first calcined in nitrogen at 700 0C. The gas was then switched to pure CO2 and the weight gain continuously tracked. After reaching the ultimate conversion, the gas was switched back to N2. This process was repeated for 2-3 cycles. The data obtained on Aldrich CaCO3 and PCC undergoing this cyclical study is shown in Figure 10. It can be seen that the reactivity of Aldrich CaCO3 exhibited a gradual decrease even after the first cycle. In contrast, PCC completely regained its mass after the first calcination and carbonation cycle. At 700 0C, we can deduce that the conversion is almost complete (>95%). The figure also shows that the reactivity did not decrease in the second cycle either. Under the reaction conditions chosen, any sintering did not seem to adversely affect the sorbent morphology. We continued an extended study of eleven calcination and carbonation cycles lasting over three days on PCC. The data is provided in Figure 11. It can be seen that the sorbent reactivity remained high and if enough reaction time is provided, the conversion could reach beyond 90% in every cycle. This is a positive result for the structural viability of this sorbent under multiple cycles. Effect of Vacuum Calcination
[0094] The effect of initial surface area of CaO sorbents was studied. CaO sorbents were synthesized from PCC under different calcination conditions. The role of surface area on the extent of carbonation is shown in Figure 12. Different surface area PCC-CaO sorbents were synthesized by the calcination of PCC at a range of calcination temperature to induce varying degrees of sintering. It can be seen that a higher initial surface area (and its associated pore volume) leads to higher reactivity and conversion. Thus, it is necessary to identify calcination conditions that optimize the SA/PV and pore size distribution of PCC-CaO. It has been suggested in literature that CaO procured from the calcination of limestone under vacuum has a higher reactivity. It was observed that under air calcination at 650 - 800 0C, sharp edges of calcite powder were replaced by rounded surfaces and neck areas indicating severe sintering (Beruto, D., and Searcy, A.W., "Calcium oxides of high reactivity." Nature, 1976, 263, 221-222). The resulting CaO structure was highly crystalline as well. In contrast, the sharp edges of calcite were retained in the CaO obtained under vacuum. The CaO however did not possess a high degree of crystallinity. The latter also showed high reactivity towards hydration. Vacuum calcination leads to the formation of a metastable-nanocrystalline calcia structure while calcination in helium atmosphere lead to a stable microcrystalline calcia structure (Dash, S., Kamruddin, M., Ajikumar, P.K., Tyagi, A.K., and Raj, B., "Nanocrystalline and metastable phase formation in vacuum thermal decomposition of calcium carbonate." Thermochimica acta, 2000, 363, 129-135). Beruto et al., [1980] estimated the surface area and pore volume of limestone based CaO to be about 78-89 m2/g and 0.269 ml/g respectively. [0095] The effect of vacuum calcination was studied in this process. The surface area of Linwood carbonate increased from 17.79 to 21.93 m2/g and pore volume from 0.07815 to 0.1117 ml/g for calcination under nitrogen and under vacuum, respectively. Similar enhancements were observed for PCC based CaO sorbents as well. It has been observed that PCC-CaO is susceptible to high degree of sintering and the surface area of the sorbent falls off rapidly. Calcination in nitrogen resulted in surface areas below 13 m2/g repeatedly. However, vacuum calcination lead to a surface area of 19.84 m2/g and 0.04089 ml/g pore volume. The carbonation characteristics are shown in Figure 13.
[0096] Vacuum calcination of PCC followed by the carbonation of PCC-CaO was repeated over two cycles. PCC was first vacuum calcined to CaO-1 at 750 0C. CaO-1 was carbonated to CC-2 at 700 0C followed by its vacuum decomposition to CaO-2 that is carbonated to CC-3. The values of surface area and pore volume of the sorbent at various stages are provided in Table 3 below:
Figure imgf000033_0001
Table 3: Structural properties of Calcium based sorbents undergoing vacuum calcination at 750 0C and carbonation at 7000C.
[0097] The data shows that PCC is susceptible to sintering because the CaO obtained in the first cycle has a surface area of only 12.63 m2/g compared to 38.3 m2/g of PCC. As expected, pore filling leads to a drop in both properties when CaO 1 carbonates. The extent of carbonation was beyond 90%. However, it can be seen that the SA of CaO obtained after the second vacuum calcination step, CaO 2, is 15.93 m2/g, which is higher than the SA of CaO 1. The pore volume of CaO 2 is also higher than that of CaO 1. These results prove that there is no systematic decline in SA and PV of sorbents with increasing calcination-carbonation cycles and that this combination is capable of providing a sustained conversion over many cycles. [0098] The article "Carbonation-Calcination Cycle Using High Reactivity
Calcium Oxide for Carbon Dioxide Separation from Flue Gas" by Himanshu Gupta and Liang-S. Fan, published on the web July 11 , 2002 by Ind. Eng. Chem. Res. 2002, 41 , 4035-4042 is hereby incorporated in its entirety by reference. Enhanced Hydrogen Production Integrated with CO2
Separation in a Single-Stage Reactor
[0099] A variety of chemical processes known to generate syngas include:
Steam Gasification: C + H2O → CO + H2 (X)
Steam Methane Reforming: CH4 + H2O → CO +3H2 (X) Partial oxidation of Hydrocarbon: CxHy +O2 → CO + H2 (X)
[00100] The flow sheet shown in Figure 14 integrates the Calcium-based reactive separation process under development in this project with a coal gasifier based electric power/chemical synthesis process plant 140. The main coal gasifier 140a consists of a high pressure and high temperature unit that allows contact between coal 140b, steam 14Oe and air/pure oxygen 14Oy in a variety of schemes. Boiler feed water 14Od is preheated by passing it through gasifier 140a prior to steam tubine 140c. Waste from the gasifier is collected as slag 14Oz. Typical fuel gas compositions from various known coal gasifiers are shown in Table 4. Once the water gas mixture is formed at the exit of the gasifier 140a, CaO fines are injected 14Of into the gas duct that react with the CO2 present in the gas mixture leading to the formation of solid CaCO3. As the fuel gas flows past the WGS catalyst monoliths 14Og, the WGS reaction is effected forming more CO2 in the process. The entrained CaO particles react with the incipiently formed CO2 gas, thereby allowing further catalysis of the WGS reaction to occur. This process can be tailored to attain as high a H2 concentration as possible. At the exit of the WGS reactor, the reacted CaCO3 particles are captured using a high temperature solids separator 14Oh (e.g., a candle filter or a high temperature ESP) and separated fuel gas stream. The spent solids are now sent to a rotary calciner 140k to thermally decompose the CaCO3 14Oj back to CaO 14Of and pure CO2 140m. The high purity CO2 gas can now be economically compressed 1401, cooled, liquefied and transported for its safe sequestration 140m. The rotary calciner allows the calcium particles to remain segregated, which is crucial in maintaining a sorbent structure characterized by a higher porosity. It was previously observed in our studies that heaping of calcium sorbents leads to a lower porosity and consequently a lower reactivity over the next carbonation cycle. The calcination of the sorbent can also be effected under sub-atmospheric conditions that allow the removal of CO2 as soon as it is formed from the vicinity of the calcining sorbent, thereby aiding further calcination. This vacuum can be created by means of ejector systems that are widely used in maintaining vacuum in large vacuum distillation units (VDU) in the petroleum refining industry. Lock and hopper combinations and appropriate seals ensure that the sorbent can be effectively separated from the CO2 stream and re-entrained in the fuel gas duct. The hydrogen enriched fuel gas 14Oi can now be used to generate electric power in a fuel cell 14On or used to make fuels and chemicals 14Oq without any low temperature clean up. The fuel cell may receive a supply of air 14Op and discharge steam 140o. The hydrogen enriched fuel gas may be sent to gas turbine 14Or used to drive generator 14Ot to produce electricity and air compressor 140s to produce a stream of compressed air. The stream of compressed air may be sent to air separator 14Ox to produce the air/oxygen of 14Oy. The discharge from gas turbine 14Ot may be sent through heat exchanger 14Ou prior to being discharged at stack 14Ov. The absorbed heat may be collected by steam turbine 14Ow to produce additional electricity. Thermodynamic analysis
[00101] Primarily three important gas-solid reactions can occur when calcium oxide (CaO) is exposed to a fuel gas mixture obtained from coal gasification. CaO can undergo hydration, carbonation and sulfidation reactions with H2O, CO2 and H2S, respectively. These can be stoichiometrically represented as: Hydration: CaO + H2O } Ca(OH)2 (5)
Carbonation: CaO + CO2 _> CaCO3 (6)
Sulfidation: CaO + H2S * CaS + H2O (7) [00102] All these reactions are reversible and the extent of each of these reactions depends on the concentrations of the respective gas species and the reaction temperature. Detailed thermodynamic calculations were performed to obtain equilibrium curves for the partial pressures of H2O (PH2O), CO2 (PCO2) and H2S (PH2S) as a function of temperature, for the hydration, carbonation, and sulfidation reactions using HSC Chemistry v 5.0 (Outokumpu Research Oy, Finland). The equilibrium calculations were based on the fuel gas compositions that are typical of the different types of coal gasifiers. The details of the fuel gas mixtures are illustrated in Table 4.
Table 4: Typical fuel gas compositions obtained from different gasifiers. (Stultz and Kitto, 1992)
Figure imgf000036_0001
Figure imgf000037_0001
[00103] The relationship between the reaction temperature and the equilibrium partial pressures of H2O and CO2 for the hydration and carbonation reactions are shown in Figure 15 (a). For a typical gasifier moisture composition ranging from 12-
O 20 atm (PH2O) the hydration of CaO occurs for all temperatures below 550-575 C, respectively. By operating above these temperatures, the CaO-hydration can be prevented. Figure 15 (b) shows the typical equilibrium CO2 partial pressures (PCO2) as a function of temperature. From the data in Table 4, it can be inferred that the typical PCO2 in the gasifiers ranges from 0.4-4.3 atm for entrained flow (slurry) and entrained flow (dry) gasifier systems respectively. The equilibrium temperatures
O corresponding to those PCO2 lie in the 830-1000 C range as shown in Figure 15(b). Thus, by operating below these temperatures, we can effect the carbonation of CaO. For the reversible sulfidation of CaO (eqn 7) the thermodynamic calculations depend on the concentration of moisture in the system. Hence, Figure 16 depicts the equilibrium H2S concentrations in ppm for varying moisture concentrations (PH2O)
O and 30 atm total pressure. For a typical operating temperature range of 800-1000 C the equilibrium H2S concentration is between 5700-1700 ppm respectively for 20 atm
0
PH2O. Consequently, at 800 C we need more than 5700 ppm H2S for the sulfidation o of CaO to occur. This number changes to 570 ppm for a PH2O of 2 atm at 800 C. Thus, by changing the moisture/steam concentration in the system we can prevent the sulfidation of CaO from occurring.
EXPERIMENTAL Sorbent and Catalyst Characterization
[00104] The high and low temperature water gas shift (WGS) reaction catalysts were procured from Sϋd-Chemie Inc., Louisville, KY. The high temperature shift (HTS) catalysts comprises of iron (III) oxide supported on chromium oxide. Precipitated calcium carbonate (PCC) was synthesized by bubbling CO2 through a slurry of hydrated lime. The neutralization of the positive surface charges on the CaCO3 nuclei by negatively charged N40V® molecules forms CaCO3 particles characterized by a higher surface area/pore volume and a predominantly mesoporous structure. Details of this synthesis procedure have been reported elsewhere (Agnihotri et al., 1999). Hydrated lime from a naturally occurring limestone (Linwood Hydrate, LH) and a naturally occurring limestone (Linwood Carbonate, LC) was obtained from Linwood Mining and Minerals Co.
[0101] The sorbents and catalyst were analyzed to determine their morphologies using a BET analyzer. The BET surface areas, pore volumes, and pore size distributions of the catalysts and sorbents were measured at -196 0C using nitrogen as the adsorbent in a Nova 2200 Quantachrome BET analyzer. Special care was taken to ensure that all samples were vacuum degassed at 250 0C for 5 hours prior to BET analysis.
WGS Reactor Setup [0102] A reactor setup was designed, underwent several iterations and was assembled to carry out water gas shift reactions in the presence of CaO and catalyst. The reactor design assembly used to carry out these experiments is shown in Figure 17. This setup enables us to carry out both the water gas shift reaction in the presence of CaO as well as the regeneration of the sorbent in flowing gas such as nitrogen and/or steam. The setup 170 consists of a tube furnace 17Op, a steel tube reactor 170a, a steam generating unit 170c, a set of gas analyzers for the online monitoring of CO and CO2 concentrations 17On, a condenser 170m to remove water from the exit gas stream and a high pressure water syringe pump 170b. [0103] All the reactant gases (H2, CO, CO2, and N2) are metered using modified variable area flowmeters 17Oe - h respectively. The syringe pump is used to supply very accurate flow-rates of water into the heated zone of the steam- generating unit in the 0.01-0.5 ml/min range. Once the steam is generated, it is picked up by the CO/N2 gas mixture 17Oi and enters the main reactor where the sorbent/ catalyst mixture 17Oo is loaded. All the lines connecting the steam- generating unit to the main reactor are heated using heating tapes. The steam generator is also packed with quartz wool 17Od in order to distribute the water drops as they enter into the heating zone. The packing is utilized in order to provide greater surface for water evaporation and to dampen out fluctuations in steam formation. The main problem with a fluctuating steam supply is that the gas analyzers used to measure the exit CO and CO2 concentrations are sensitive to gas flow rates. Even though the steam is being condensed out before the gas is sent into the analyzers, surges in the steam supply still affect the overall gas flow rate, causing the CO and CO2 readings to fluctuate. The packing ultimately ensures a more continuous and constant overall gas flow rate into the main reactor and into the analyzers. Thermocouple 170k is used to monitor the temperature inside reactor 170a. Any extra gas inlets of reactor 170a are blocked 170I. [0104] A steel tube reactor is used to hold the Ca-based sorbent and catalyst, and is kept heated using a tube furnace. The sorbent loading unit of the reactor is detachable which enables easy removal and loading of the sorbent and therefore minimizes the sorbent loading time between runs. Also, the sorbent can be changed without having to cool down the entire reactor. The gas mixture 17Oj entering the reactor is preheated to the reaction temperature before contacting the sorbent/ catalyst particles. The gases exiting the reactor first flow through a condenser in order to separate out the moisture and then to a set of gas analyzers. Sub Atmospheric Calcination Reactor Setup [0105] Once the Calcium based sorbent has reacted with the CO2 being produced by the WGSR, the sorbent has to be regenerated for further use in subsequent cycles. During the regeneration of the sorbent, carbon dioxide is released from the sorbent. In order to minimize the necessity for further treatment of this released CO2 before sending it to sequestration sites, it is necessary to regenerate the sorbent such that a pure stream of CO2 is released. Vacuum calcination provides one method for ensuring that concentrated streams of CO2 are release in the regeneration phase. The detailed setup is shown in Figure 18. This setup 180 was assembled to handle the regeneration of large quantities of sorbent (~10-20g per batch). The setup includes an alumina tube reactor 180b, which would hold the sorbent samples in a split tube furnace 180c that provides the heat necessary to calcine the sorbent 18Od, two Non Dispersive Infra Red (NDIR) analyzers 180k -I to monitor the CO2 concentration (ranges 0-2500ppm and 0-20%) and two vacuum pumps 18Of and 18Oi. 10g of sorbent yields about 2.4L of CO2 at atmospheric pressure and temperature over the entire decomposition process. This gas needs to be diluted with air in order to ensure that the CO2 concentration lies in the detection range of the CO2 analyzers. Vacuum Pump 18Of is a dry vacuum pump procured from BOC Edwards capable of achieving vacuum levels as low as 50mtorr and gas flowrates of 6m3/hr. The CO2 analyzers have their own inbuilt pumps and are capable of drawing up to 2LPM from the header for online CO2 analysis. The second pump 18Oi is a smaller dry pump and is put in place to ensure that there is no pressure buildup in the !4" lines connecting the vacuum pump to the analyzers. Pump 18Oi discharges to vent 18Oj. The temperature of the furnace is controlled with a thermocouple inserted into the central zone of the furnace. The temperature of the reactor was also monitored using a second thermocouple inserted into the center of the alumina tube. The setup is also capable of combining vacuum calcination with flow of sweep gas 180a. As it may not be feasible to supply very low vacuum levels for the calcination of the sorbent in industrial settings, it may be necessary to study the calcination process in combination with the addition of various sweep gases such as N2/ steam. Pressure gauges 18Oe, h and volumetric flow meters are included to monitor the vacuum pressure in the reactor, the pressure in the VA" lines and the flows of the sweep gases into the calciner and the flow of the air 18Og used in the dilution of the exhaust CO2 before sending it to the analyzers. The analyzers are also connected to a data acquisition system 180m that can record analyzer readings every second. RESULTS AND DISCUSSIONS
Catalyst and Sorbent Characterization
[0106] The characterization of the high temperature shift (HTS) catalyst in a
2 BET analyzer revealed that the catalyst has a BET surface area of 85 m /g and a total pore volume of about 0.3 cc/g. The majority of the pores were found to occur around 180 A as evident from the maximum in its pore size distribution plot shown in Figure 19. In contrast, the low temperature shift (LTS) catalyst has a BET surface
2 area of 52 m /g and a total pore volume of about 0.1 cc/g. The majority of these pores were found to occur around 37 A as evident from the maximum in its pore size distribution plot (Figure 19). [0107] The surface area (SA) and pore volume (PV) of the three different CaO precursors are provided in Table 5. Figure 20 shows the pore size distribution (PSD) of these precursor fines. It can be seen that LC fines do not have high SA/PV.
However, upon calcination and subsequent hydration, the SA/PV of the calcium hydroxide (LH) fines increase as can be observed for the LH sample. The porosity is maximized in the microporous range (30-50 A range). In contrast, the SA/PV of the morphologically altered PCC are much higher. Further, most of the porosity lies in the 100-300 A range.
Table 5: Morphological properties of the natural and synthesized CaO precursors and the HTS catalyst obtained from BET analyses.
Figure imgf000042_0001
Water gas shift reaction (WGSR): Catalyst Testing and Analysis
[0108] The HTS catalyst was tested for its catalyst activity towards the WGS reaction between 500-700 0C. Blank runs (without any sorbent) were performed in a reaction mixture comprising of 3% CO and 9% H2O, the balance being 5.0 grade N2. The total gas flow-rate was maintained at about 1.5 slpm and the steam/CO ratio was set at ~3. Typically about 0.5 grams of the HTS catalyst was loaded in the reactor prior to each run. The catalyst activity increases monotonically with increasing reaction temperature. This is evident from Figure 21 below. The CO conversion increases from 24.3% at 500 0C to 69.3% at 550 0C. It finally reaches about 80% at 600 0C. Beyond 600 0C the conversion does not change much but remains steady at -78% at 700 0C. This might be due to the equilibrium limitations governing the WGS reaction scheme is depicted in eqn (8) below:
CO + H2O → CO2 + H2 (8)
The data were further analyzed to check if the system was operating within the domain of WGS equilibrium. The thermodynamic equilibrium constant (K) for any temperature for this reaction was computed using the software "HSC Chemistry v 5.0" (Outokumpu Research Oy, Finland). The observed ratio was computed from the experimental data by obtaining the ratio of the partial pressures of the products and the reactants as per the eqn (9) below:
V Kobs - (9)
\rH2 )\rCO2 ) Figure 22 illustrates the effect of temperature on the ratio of partial pressures (Kobs) obtained from the experimental data. This is compared with the thermodynamic equilibrium values (Keq). From the figure it is evident that we are operating in the region that is below the thermodynamic equilibrium. At 500 °C the Kobs is 0.028 while the corresponding Keq is 4.77. Keq monotonically decreases with increasing temperature. In contrast, Kobs increases with temperature for our operating conditions. Thus, at 600 0C the Kobs increases to 1.4 while the Keq moves down to 2.5. This trend continues and it is clearly evident from the figure that the system moves closer to equilibrium as we progressively increase the temperature from 500 to 700 0C. Combined Carbonation and Water gas shift reaction: Sorbent Testing and Analyses
[0109] The combined carbonation and WGS reaction was tested in the reactor assembly used for the catalyst testing. The experimental conditions were exactly identical to that used for testing the catalyst. The runs were performed in a reaction mixture comprising of 3% CO and 9% H2O, the balance being 5.0 grade N2. The total gas flow-rate was maintained at about 1.5 slpm and the steam/CO ratio was set at ~3. Typically about 0.5-1 g of the HTS catalyst was loaded in the reactor prior to each run. Different calcium oxide precursors were tested. Naturally occurring limestone, Linwood Carbonate (LC) and the corresponding hydrated lime, Linwood Hydroxide (LH) were obtained from Linwood Mining and Minerals Co. The structurally modified calcium carbonate (PCC) was prepared in-house and the details are outlined below.
Sorbent testing without catalyst [0110] The sorbents were initially tested for catalytic activity towards WGSR and CO conversion without any HTS catalyst from 500-700 0C. This would obviate the need for any catalyst in the system. However, detailed investigation resulted in very miniscule activity and hence it was concluded that HTS catalyst was required for further combined reaction testing. Combined reactions with PCC-HTS catalyst system
[0111] Typically about 0.5 g of HTS catalyst and 1.5 g of PCC were loaded in the reactor and the temperature was ramped till 700 0C in flowing N2. This procedure ensured the calcination of the calcium carbonate to calcium oxide and it was monitored using CO2 analyzer. Subsequently, the reaction temperature was lowered to 600 0C and the reaction gas mixture was allowed to flow through the system. The CO analyzer continuously monitored the CO flow through the system and the breakthrough curve depicting the CO conversion with time is as shown in Figure 23 below. The system gives almost 100% conversion for first 240 seconds (4 min) following which the initial reactivity of the sorbent slowly falls to give about 90% CO conversion at 1000 seconds (16.5 min). The sorbent gradually achieves its maximum loading capacity with time and finally at around 2500 seconds (42 min) the sorbent reaches its breakthrough loading. Beyond this the CO conversion of 81 % corresponds to that obtained with only the catalyst at 600 0C. This can be validated from Figure 21. [0112] The system was then switched to pure N2 flow and the reaction temperature was increased to 700 0C to drive the calcination of the CaCO3 formed due to carbonation. Thus the reactions occurring in the system are: Reaction phase:
WGSR: CO + H2O → CO2 + H2 (7)
Carbonation: CaO + CO2 → CaCO3 (8)
Regeneration phase:
Calcination: CaCO3 → CaO + CO2 (9)
The termination of the calcination was ensured by monitoring the CO2 released using a CO2 analyzer. The reaction temperature was again lowered to 600 0C and the sorbent-catalyst system was subjected to the reaction mixture for a second reaction cycle. The 2nd cycle CO breakthrough curve is also depicted in Figure 23. It is evident from the figure that the CO conversion is not as superior as in the 1st cycle. The CO conversion monotonically decreases to about 90% in 110 seconds, 80% in 240 seconds and gradually to about 50%. It is interesting to note that at the end of the breakthrough the sorbent-free catalytic CO conversion of 81% is not achievable. This could be attributed to the loss in the catalytic activity after the first regeneration cycle. This is because the catalyst is subjected to CO2, an oxidizing atmosphere, during the calcination phase. Thus the deactivated catalyst is not able to augment the WGS reaction kinetics and hence we see a poor performance of the sorbent-catalyst system in the 2nd cycle. The solitary sorbent has been subjected to numerous carbonation calcination cycles and has shown satisfactory performance (Iyer et al, 2004).
Combined reactions with LH-HTS catalyst system
[0113] Typically about 1 g of the HTS catalyst and 1.3 g of LH were loaded in the reactor and the temperature was ramped up slowly till 600 0C in flowing N2. This procedure ensured the calcination of the calcium hydroxide to calcium oxide. Calcium hydroxide decomposes above 400 0C. Subsequently, the reaction gas mixture was allowed to flow through the system. The CO analyzer continuously monitored the CO flow through the system and the breakthrough curve depicting the CO conversion with time is as shown in Figure 24 below. The system gives almost 100% conversion initially to give about 90% CO conversion at 900 seconds (15 min). The sorbent gradually achieves its maximum loading capacity with time and finally at around 3000 seconds (50 min) the sorbent has achieved its breakthrough loading. Beyond this the CO conversion of 81% corresponds to that obtained with only the catalyst at 600 0C as was shown in Figure 21. [0114] The system was then switched to pure N2 flow and the reaction temperature was increased to 700 0C to drive the calcination of the CaCO3 formed due to carbonation. Subsequently, the reaction temperature was lowered to 600 0C and the LH-CaO/catalyst system was subjected to the reaction mixture for a second reaction cycle. The 2nd cycle CO breakthrough curve is also depicted in Figure 24. It is evident from the figure that the CO conversion is not as superior as in the 1 st cycle. The CO conversion monotonically decreases to about 90% in 60 seconds, 80% in 180 seconds and gradually to about 30%. It is interesting to note that at the end of the breakthrough the sorbent-free catalytic CO conversion of 81% is not achievable. This could be attributed to the loss in the catalytic activity after the first regeneration cycle. This is because the catalyst is subjected to CO2, an oxidizing atmosphere, during the calcination phase. Thus the deactivated catalyst is not able to augment the WGS reaction kinetics and hence we see a poor performance of the sorbent- catalyst system in the 2nd cycle. The solitary sorbent had been subjected to numerous carbonation calcination cycles and has shown satisfactory performance over few cycles.
Comparison of the PCC and LH sorbents
[0115] Figure 25 compares the CO conversion breakthrough curves for the
PCC and LH sorbent-catalyst systems. The curves are for the 1st reaction cycle. The CO conversion at any given time for PCC-CaO is always higher than that of LH-CaO. The PCC system gives almost 100% conversion for first 240 seconds (4 min) while the LH sorbent system sustains this conversion only in the initial few seconds. Subsequently, the PCC system gives about 90% CO conversion at 1000 seconds (16.5 min) followed by 85% in 1600 seconds (27 min). In contrast, the LH system gradually gives about 90% CO conversion at 900 seconds (15 min) and followed by 85% in 1200 seconds (20 min). Both the sorbent systems gradually achieve their maximum loading capacity with time and finally at around 2500-3000 seconds they reach their breakthrough loading. Beyond this the CO conversion of 81% corresponds to that obtained with only the catalyst at 600 0C. Hence, it is evident from Figure 24 that the PCC-CaO performance dominates over that of LH-CaO at any given time. [0116] Figure 26 illustrates the generation 1 MWe of steam.
[0117] Figure 27 illustrates one embodiment of the present invention providing
1.002 MWe total capacity.
[0118] Figure 28 illustrates a second embodiment of the present invention providing 1 MWe total capacity.
[0119] Figure 29 illustrates another embodiment of the present invention providing 1.33 MWe total capacity.
[0120] Figure 30 illustrates yet another embodiment of the present invention providing 1.33 MWe total capacity [0121] Figure 31 illustrates an alternative embodiment of the present invention providing 1.54 MWe total capacity.
[0122] Figure 32 illustrates yet another alternative embodiment of the present invention providing 1.07 MWe total capacity.
[0123] Figure 33 illustrates an alternative embodiment of the present invention providing 1 MWe total capacity.
[0124] Figure 34 illustrates an alternative embodiment of the present invention providing 1 MWe total capacity.
[0125] Figure 35 illustrates yet another embodiment of the present invention providing 1.54 MWe total capacity. [0126] Figure 36 illustrates an alternative embodiment'of the present invention providing 1 MWe total capacity at 80% CO2 capture.
[0127] Figure 37 illustrates another embodiment of the present invention providing 300 MWe total capacity at 90 CO2 capture.
REGENERATION OF CALCIUM SULFIDE
[0128] Process Objective: [0129] Many processes that pertain to the use of calcium based sorbents
(CBS) for the control of acid gases such as H2S, CO2 and S02 from combustion and gasification of fossil fuels such as coal, oil and natural gas are hampered by the incomplete reactivity of calcium towards various gas-solid reactions such as sulfation to form CaSO4, sulfidation to form CaS, and carbonation to form CaCO3. High conversion of CBS is essential in lowering the levelized operating cost of the concerned process. One of the methods to enhance reactivity of these CBS is to recycle the unreacted sorbent back into the flue and fuel gas mixtures. [0130] SO2 Capture: [0131] The proposed methodology involves the injection of inexpensive limestone, which cost about $l 0-20/ton, into the upper furnace region of a fossil fuel boiler where it simultaneously decomposes to CaO and forms CaSO4 at the high temperature. The solids traverse the flue gas duct and ultimately the particulate capture devices capture a well-mixed solid sorbent mixture consisting of fly ash, CaO and CaSO4.
[0132] This process involves the extraction of the unreacted calcium (in the form of CaO and Ca(OH)2) from the solid mixture by an aqueous sugar solution that forms complexes with calcium. Other chemicals that can achieve this complexation include OH-containing groups such as phenols, glycerol, glycols, EDTA, etc. [0133] This complexation leads to the selective removal of only unreacted calcium from1 the solid mixture by forming water-soluble complexes. The slurry is then filtered and the residue now contains fly ash and calcium sulfate (which signifies 100% conversion of calcium). The calcium escapes with the filtrate. The solution is taken into another vessel where it is subjected to carbonation in which gaseous C02 is injected into the sugar solution containing calcium in the form of chemical complexes. The carbonation reaction leads to the formation of calcium carbonate crystals that can be removed by filtration. The CaC03 is now dried and injected for further SO2 capture in the FSI mode. The drop in weight of the sorbent, as shown in Figure 38 to 56% in the calcination phase indicates the formation of pure CaCO3 only. This also proves that other materials, such as fly ash and CaSO4 do not leach out into the sugar solution and mix with the precipitated CaCO3. [0134] The advantages are:
[0135] 1. The process is based on inexpensive naturally occurring limestone.
[0136] 2. The process DOES NOT necessitate high sorbent conversion to CaSO4 [0137] 3. However, as seen in Figure 38, this CaCO3 attains almost 87.5% conversion to sulfation reaction, which is much higher than the conversion of natural limestone, which is of the order of 30^40%.
[0138] 4. The process obviates the use of expensive organic surface charge modifying dispersants, which also enhance sorbent reactivity. [0139] C02 capture from Flue/Fuel Gas:
[0140] A CaO based CO2 separation process is under development and patent protection at OSU. It involves the injection of CaO fines into flue/fuel gas mixtures to reactively separate CO2 in the form of CaCO3. These particles are physically removed from the gas duct and separately calcined to release pure C02 stream and CaO, which are used in the next carbonation phase. This cyclic carbonation and calcination reaction (CCR) process leads to CO2 separation.
[0141] However, H2S and SO2 attacks the CBS forming CaS and CaSO4 respectively, which reduces the sorbent capacity for C02 capture. The complexation-based process can be used to obviate the challenge that H2S and SO2 pose. The mixture containing the sorbents is treated with complexing agent solution (such as sugar solution) to extract out unreacted CaO as explained above. The Calcium carbonate so formed is then sent to the catchier to convert it to CaO. This fresh and pure CaO is then re-introduced into the CCR scheme.
SELECTIVE RECLAIMATION
[0142] The foreseeable enhanced use of gasification systems is based upon their higher energy to electricity efficiency, ability to convert fossil fuel energy to hydrogen, and an envisaged ease of carbon management thereby increasing their importance in a carbon constrained world. Numerous agencies are promoting the use of coal, which occurs abundantly, in gasification systems to aid hydrogen production for the envisaged hydrogen economy and to enhance domestic energy security. However, sulfur present in coal leads to the formation of hydrogen sulfide (H2S) under the reducing conditions in the gasifier. Downstream requirements imposed by the gas turbine and the purity standards of hydrogen for its use in chemical production and PPM fuel cell operation necessitate that the H2S be reduced significantly. Based on the exact requirement, the concentration of H2S could be as low as 10 ppm. [0143] One of the processes that reduce H2S involves its reaction with calcium-based sorbents such as CaO and CaCO3 to forms calcium sulfide (CaS) through a heterogeneous, high temperature gas solid reaction (Fenoil and Lynn, 1995 a,b; Chauk et al., 2000). However CaS cannot be safely disposed as it can evolve H2S gas or cause sulfide ions to leach into ground water. Hence, CaS needs additional chemical treatments to stabilize it. There are two approaches to handling CaS. In one scheme, high temperature reactions that completely convert CaS back to CaCO3 can be employed by its reaction with steam and CO2 (Adanez, et al., 2001). This reaction, carried out at 650°C, can be written as: CaS + H2O + CO2 → CaCO3 +H2S
[0144] This process provides a stream of H2S as well as CaCO3 that can be recycled back for additional H2S removal in a subsequent cycle. However, this leads to a loss in sorbent reactivity after a few cycles of sulfidation regeneration reaction. The sulfidation extent of CaCO3 dropped from 69% in the first cycle to 13% at the end of 21 cycles (Keaims et al., 1974). Another process involves the high temperature oxidation (at 920°C) of CaS to form CaSO4, which is more amenable to land filling, as it is a stable and relatively unreactive compound (Qiu, et al., 1999). However, the formation of a larger molar volume product (CaSO4) leads to pore pluggage and pore mouth closure, which limit the sorbent from attaining complete conversion. Hence it is possible that the sorbent always contained unconverted CaS, thereby making this oxidation procedure difficult to deploy in practice.
[0145] A different approach can be employed that converts CaS back to
CaCO3 in an aqueous phase. The regeneration of CaS can be carried out by bubbling CO2 through aqueous slurry of CaS. The overall stoichiometric reaction can be written as: CaS + H2O + CO2 → CaCO3 + H2S [0146] This reaction, which occurs at ambient temperature, goes to completion and leads to the formation of H2S. H2S laden gas can be processed using conventional processes such as the Claus process to convert it to sulfur or sulfuric acid, which can add revenue streams to the entire plant operations. Currently, a low temperature amine based process that separates H2S from fuel gas originating from coal gasification does indeed create revenue stream at the Tampa Electric's Polk Power Plant in Florida by selling sulfuric acid to the fertilizer industry in the vicinity. The use of a regenerative calcium based desulfurization process has additional advantages. For example, the use of calcium over multiple cycles drastically reduces the requirement of fresh calcium sorbent, as a majority of it is recycled. A related benefit is that the quantity of byproducts that require additional management such as land filling is also reduced.
[0147] LITERATURE REVIEW
[0148] A few studies have been reported in the literature on this reaction.
Biswas et al. (1976) aimed at sulfur recovery from gypsum. They studies on the aqueous phase carbonation of reduced gypsum confirmed that a high degree of conversion could be attained by this process. In fact less than 1 % sulfur was left after the carbonation reaction. The hypothesized this reaction to occur in two steps:
[0149] 2 CaS + H2O + CO2 → Ca(HS)2 + CaCO3
[0150] Ca(HS)2 + H2O + CO2 → 2 H2S + CaCO3 [0151] This aqueous phase reaction exhibits a sigmoidal conversion versus time plot. The reaction rate is slow in the beginning due to the limited solubility of
CaS in water. However, H2S formed as a result of carbonation, starts reacting with the
CaS itself converting it to higher solubility Ca(HS)2. Eventually the carbonation slows down due to the depletion of Ca(HS)2 in the slurry. [0152] Jia and Lu (1992) continued this work by attempting to derive rate constants for the carbonation of CaS. They observed a slight optimization in the extent of conversion at 60°C. This optimization is a balance between enhanced reaction rates and a decrease in absorption of CO2 in water and CaS solubility as the temperature increases. They estimated that the concentration of H2S in the exit gases (0.13 to 0.96) is sufficient for a Claus process following this carbonation step. A higher solid CaS loading in the slurry hastened the reaction due to an inherent rise in temperature as the reaction proceeds.
[0153] Nishev and Pelovski (1993) detailed the effect of reaction temperature and PCO2 on the kinetics of carbonation of CaS. An increase in PCo2 increases the rate of the reaction as well as the conversion. A higher PCo2 leads to faster nucleation that leads to a less porous CaCO3 sorbent. Internal diffusion control dominated the progress of the reaction as evidenced by the activation energy values in the 17-27 kJ/mol. It was also shown that a higher Pco2 lead to a higher value of activation energy and pre-exponential factor. [0154] Brooks and Lynn (1997) investigated the recovery of calcium carbonate and hydrogen sulfide from spent calcium sorbents. They used a methyldiethanolamine (MDEA) mediated carbonation procedure that enhances the reaction rate by creating complexes with both CO2 and H2S. The reactions are: [0155] MDEA-H2S + CaS → MDEA + Ca(HS)2 [0156] MDEA-H2CO3 + MDEA + Ca(HS)2 → 2 MDEA. H2S + CaCO3
[0157] In this project we are aiming to recover CaS to form H2S and CaCO3.
This aqueous phase carbonation provides a method to completely regenerate calcium- based sorbents, which lose reactivity towards high temperature reactions that are carried out in a cyclic fashion. Further it is our aim to study the morphology of the CaCO3 that is formed for its surface area and pore volume characteristics, which are crucial for sorbent reactivity. A prior study has already indicated the efficacy of OSlTs patented mesoporous sorbent towards the extent of sulfidation compared to naturally occurring limestone (Chauk et al., 2000). The main aim of this portion of the project is to recreate the mesoporous PCC structure starting from CaS (pure and CaS resulting from sulfidation of CaO/CaCO3) instead of the conventional Ca(OH)2 starting material.
[0158] The following sections detail our efforts in this direction to date.
[0159] EXPERIMENTAL RESULTS AND DISCUSSION
[0160] Chemical Analysis of CaS: Numerous analytical techniques including iodine-based titration were tried to accurately quantify sulfide content. However, the variability in the analytical procedure warranted another technique for its analysis. We have eventually developed a procedure that first involves the liberation of H2S from CaS by the addition of HCI acid. This mixture is heated in refluxing mode to ensure the completion of this reaction. The exiting gases are then sent through a gas sparger into a solution of NaOH and H2O2. The NaOH traps the H2S in the form of Na2S and the H2O2 stabilized it further by converting it to sulfate ions, SO-I 2'. The concentration of sulfate ions is estimated using an Ion Chromatography setup procured from Alltech. A mobile phase consisting of NaHCO3/Na2CO3 provides adequate peak resolution on an Allsep 7u anionic column. [0161] Liquid Phase Studies of CaS Carbonation:
[0162] XRD Results: It was difficult to evolve H2S by heating a solution of CaS in water, even to boiling. Figure 39 shows the XRD plot of the product that was obtained after drying it. A similar trend in crystal structure was observed in the product obtained even when helium is bubbled through aqueous CaS slurry. [0163] The effect of CO2 gas was studied next. Aqueous slurry is made by mixing 1 gram of CaS in 250 ml of water. Pure CO2 gas is then bubbled through the slurry containing 1 gram of CaS in 250 ml_ water as it passes through the fritted quartz distributor. The solid product is filtered and dried in a vacuum oven before subjecting it to various physical and chemical characterizations. Figure 40 and 41 reveals the XRD patterns of the solid after 10 and 60 minutes of carbonation. It can be seen that the peaks have significantly shifted away from the CaS structure, signifying the occurrence of the carbonation reaction.
[0164] Structural Properties: Preliminary experiments were carried out to understand the effect of N40V concentration on the structural characterization of the CaCO3 formed by this carbonation reaction. In the absence of N4OV, the SA/PV values were 9.28 m2/g and 0.02447 cc/g. When the CaS/N40V weight ratio is altered to 40, we observed a decrease in SA to 2.36 m2/g, but an increase in PV to 0.08276 cc/g. The pore size distribution of the two CaCO3 products is shown in Figure 42. It can be seen that the introduction of N40V increases the PV in the sorbent, which is useful in enhancing its reactivity towards sulfidation reactions.
CONCLUSIONS
[0165] The enhanced water gas shift reaction for Hb production with in-situ carbonation was studied at 600 0C using HTS catalyst and calcium sorbents. A naturally occurring calcium precursor (Linwood hydrate, LH) and a modified mesoporous Precipitated Calcium Carbonate (PCC) were used for capturing CO2 for two successive cycles. The PCC system gives almost 100% conversion for first 4 min followed by 90% at 16.5 min. In contrast, the LH sorbent system sustains 100% conversion only in the initial few seconds and gradually gives about 90% CO conversion at 15 min. Experimental evidence clearly shows that the PCC-CaO performance dominates over that of LH-CaO at any given time.
[0166] While the invention has been described in connection with what is presently considered to be the most practical and preferred embodiments, it is to be understood that the invention is not to be limited to the disclosed embodiment(s), but on the contrary, is intended to cover various modifications and equivalent arrangements included within the spirit and scope of the appended claims, which are incorporated herein by reference.

Claims

What is claimed is:
1. A method for separating carbon dioxide from a flow of gas comprising carbon dioxide, said method comprising the steps of: directing said flow of gas to a gas-solid contact reactor, said gas-solid contact reactor containing at least one sorbent, said at least one sorbent comprising at least one metal oxide; reacting said carbon dioxide with said at least one sorbent so as to remove said carbon dioxide from said flow of gas, thereby converting said at least one sorbent into spent sorbent; calcining said spent sorbent so as to liberate said carbon dioxide from said spent sorbent, thereby regenerating said sorbent; and repeating the aforementioned steps.
2. The method according to claim 1 wherein said at least one metal oxide is selected from the group consisting of: ZnO, MgO, MnO2, NiO, CuO, PbO, and CaO.
3. The method according to claim 1 wherein said spent sorbent is a metal carbonate.
4. The method according to claim 1 wherein said sorbent has a sorption capacity of at least about 70 grams of carbon dioxide per kilogram of sorbent.
5. The method according to claim 1 wherein said sorbent has a sorption capacity of at least about 300 grams of carbon dioxide per kilogram of sorbent.
6. The method according to claim 1 wherein said sorbent has substantially the same sorption capacity after calcining as said sorbent had prior to adsorbing said carbon dioxide.
7. The method according to claim 1 wherein said calcining is performed under at least partial vacuum.
8. The method according to claim 1 wherein said calcining is performed by steam.
9. A facility practicing the method according to claim 1.
10. A method for separating carbon dioxide from a flow of gas comprising carbon dioxide, said method comprising the steps of: directing said flow of gas to a first gas-solid contact reactor, said first gas-solid contact reactor containing at least one sorbent, said sorbent comprising at least one metal oxide; reacting said carbon dioxide in said flow of gas on said sorbent in said first gas-solid contact reactor so as to remove said carbon dioxide from said flow of gas; directing said flow of gas to a second gas-solid contact reactor when said sorbent in said first gas-solid contact reactor is spent thereby forming spent sorbent, said second gas-solid contact reactor containing at least one sorbent, said sorbent comprising at least one metal oxide; reacting said carbon dioxide in said flow of gas on said sorbent in said second gas-solid contact reactor so as to remove said carbon dioxide from said flow of gas; calcining said spent sorbent from said first gas-solid contact reactor so as to generate carbon dioxide and to regenerate said sorbent; directing said flow of gas to said first gas-solid contact reactor when said sorbent in said second gas-solid contact reactor is spent, thereby forming spent sorbent; and calcining said spent sorbent from said second gas-solid contact reactor so as to generate carbon dioxide and to regenerate said sorbent.
11. The method according to claim 10 wherein said calcining of said spent sorbent from said first gas-solid contact reactor is performed under at least partial vacuum.
12. The method according to claim 10 wherein said calcining of said spent sorbent from said first gas-solid contact reactor is performed by steam.
13. The method according to claim 10 wherein said calcining of said spent sorbent from said second gas-solid contact reactor is performed under at least partial vacuum.
14. The method according to claim 10 wherein said calcining of said spent sorbent from said second gas-solid contact reactor is performed by steam.
15. The method according to claim 10 wherein said at least one metal oxide is selected from the group consisting of: ZnO, MgO, Mnθ2, NiO, CuO, PbO, and CaO.
16. The method according to claim 10 wherein said sorbent has a sorption capacity of at least about 70 grams of carbon dioxide per kilogram of sorbent.
17. The method according to claim 10 wherein said sorbent has a sorption capacity of at least about 300 grams of carbon dioxide per kilogram of sorbent.
18. The method according to claim 10 wherein said sorbent has substantially the same sorption capacity after calcining as said sorbent had prior to adsorbing said carbon dioxide.
19. A facility practicing the method according to claim 10.
20. A method for regenerating a spent sorbent for carbon dioxide, said method comprising the steps of: providing a spent sorbent, said spent sorbent comprising metal carbonate; and calcining said spent sorbent so as to liberate carbon dioxide gas and so as to regenerate said spent sorbent thereby forming a sorbent comprising a metal oxide.
PCT/US2006/025493 2005-06-28 2006-06-28 Regeneration of calcium sulfide to mesoporous calcium carbonate using ionic dispersants and selective reclamation... WO2007002882A2 (en)

Applications Claiming Priority (4)

Application Number Priority Date Filing Date Title
US69470205P 2005-06-28 2005-06-28
US60/694,702 2005-06-28
US11/255,099 2005-10-20
US11/255,099 US7618606B2 (en) 2003-02-06 2005-10-20 Separation of carbon dioxide (CO2) from gas mixtures

Publications (2)

Publication Number Publication Date
WO2007002882A2 true WO2007002882A2 (en) 2007-01-04
WO2007002882A3 WO2007002882A3 (en) 2007-02-22

Family

ID=37596071

Family Applications (1)

Application Number Title Priority Date Filing Date
PCT/US2006/025493 WO2007002882A2 (en) 2005-06-28 2006-06-28 Regeneration of calcium sulfide to mesoporous calcium carbonate using ionic dispersants and selective reclamation...

Country Status (1)

Country Link
WO (1) WO2007002882A2 (en)

Cited By (10)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
WO2009137886A1 (en) * 2008-05-15 2009-11-19 Calix Limited System and method for processing flue gas
US7678351B2 (en) 2005-03-17 2010-03-16 The Ohio State University High temperature CO2 capture using engineered eggshells: a route to carbon management
US7837975B2 (en) 2006-09-25 2010-11-23 The Ohio State University High purity, high pressure hydrogen production with in-situ CO2 and sulfur capture in a single stage reactor
US8226917B2 (en) 2003-02-06 2012-07-24 The Ohio State University Separation of carbon dioxide from gas mixtures by calcium based reaction separation
CN102773006A (en) * 2012-08-17 2012-11-14 西安瑞驰节能工程有限责任公司 Device and process for cyclic capture of carbon dioxide by taking CaO as carrier
ES2401294A2 (en) * 2011-06-24 2013-04-18 Consejo Superior De Investigaciones Científicas (Csic) Device and method for the capture of c02 by cao carbonation and for maintaining sorbent activity
WO2014200357A1 (en) 2013-06-14 2014-12-18 Zeg Power As Method for sustainable energy production in a power plant comprising a solid oxide fuel cell
CN112569896A (en) * 2020-12-07 2021-03-30 华东理工大学 Calcium oxide-based bimetal composite material, preparation method and application
WO2023014397A1 (en) * 2021-08-05 2023-02-09 Infinium Technology, Llc Production and use of liquid fuel as a hydrogen and/or syngas carrier
CN116628576A (en) * 2023-07-26 2023-08-22 中南大学 Intelligent production yield monitoring method for heat carrier lime kiln

Citations (3)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
US5520894A (en) * 1992-07-02 1996-05-28 Nederlandse Organisatie Voor Toegepast-Natuurwetenschappelijk Onderzoek Tno Process for removing carbon dioxide regeneratively from gas streams
US5779464A (en) * 1996-01-10 1998-07-14 The Ohio State University Research Foundation Calcium carbonate sorbent and methods of making and using same
US5895634A (en) * 1996-05-28 1999-04-20 Mitsubishi Heavy Industries, Ltd. Desulfurization and decarbonation process

Patent Citations (3)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
US5520894A (en) * 1992-07-02 1996-05-28 Nederlandse Organisatie Voor Toegepast-Natuurwetenschappelijk Onderzoek Tno Process for removing carbon dioxide regeneratively from gas streams
US5779464A (en) * 1996-01-10 1998-07-14 The Ohio State University Research Foundation Calcium carbonate sorbent and methods of making and using same
US5895634A (en) * 1996-05-28 1999-04-20 Mitsubishi Heavy Industries, Ltd. Desulfurization and decarbonation process

Cited By (18)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
US8226917B2 (en) 2003-02-06 2012-07-24 The Ohio State University Separation of carbon dioxide from gas mixtures by calcium based reaction separation
US7678351B2 (en) 2005-03-17 2010-03-16 The Ohio State University High temperature CO2 capture using engineered eggshells: a route to carbon management
US7837975B2 (en) 2006-09-25 2010-11-23 The Ohio State University High purity, high pressure hydrogen production with in-situ CO2 and sulfur capture in a single stage reactor
US8894743B2 (en) 2008-05-15 2014-11-25 Calix Limited Method for processing flue gas
WO2009137886A1 (en) * 2008-05-15 2009-11-19 Calix Limited System and method for processing flue gas
US8632626B2 (en) 2008-05-15 2014-01-21 Calix Limited System and method for processing flue gas
AU2009246062B2 (en) * 2008-05-15 2014-07-17 Calix Limited System and method for processing flue gas
ES2401294A2 (en) * 2011-06-24 2013-04-18 Consejo Superior De Investigaciones Científicas (Csic) Device and method for the capture of c02 by cao carbonation and for maintaining sorbent activity
ES2401294R1 (en) * 2011-06-24 2013-07-11 Consejo Superior Investigacion DEVICE AND PROCEDURE FOR THE CAPTURE OF CO2 BY CARBONATION OF CAO
CN102773006A (en) * 2012-08-17 2012-11-14 西安瑞驰节能工程有限责任公司 Device and process for cyclic capture of carbon dioxide by taking CaO as carrier
WO2014200357A1 (en) 2013-06-14 2014-12-18 Zeg Power As Method for sustainable energy production in a power plant comprising a solid oxide fuel cell
EP3007802A4 (en) * 2013-06-14 2017-03-08 ZEG Power AS Method for sustainable energy production in a power plant comprising a solid oxide fuel cell
CN111482068A (en) * 2013-06-14 2020-08-04 Zeg动力股份公司 Method for sustainable energy production in a power plant comprising solid oxide fuel cells
CN112569896A (en) * 2020-12-07 2021-03-30 华东理工大学 Calcium oxide-based bimetal composite material, preparation method and application
CN112569896B (en) * 2020-12-07 2023-08-25 华东理工大学 Calcium oxide-based bimetal composite material, preparation method and application
WO2023014397A1 (en) * 2021-08-05 2023-02-09 Infinium Technology, Llc Production and use of liquid fuel as a hydrogen and/or syngas carrier
CN116628576A (en) * 2023-07-26 2023-08-22 中南大学 Intelligent production yield monitoring method for heat carrier lime kiln
CN116628576B (en) * 2023-07-26 2023-10-13 中南大学 Intelligent production yield monitoring method for heat carrier lime kiln

Also Published As

Publication number Publication date
WO2007002882A3 (en) 2007-02-22

Similar Documents

Publication Publication Date Title
US7618606B2 (en) Separation of carbon dioxide (CO2) from gas mixtures
CA2613698C (en) Separation of carbon dioxide (co2) from gas mixtures by calcium based reaction separation (cars-co2) process
US7067456B2 (en) Sorbent for separation of carbon dioxide (CO2) from gas mixtures
WO2007002882A2 (en) Regeneration of calcium sulfide to mesoporous calcium carbonate using ionic dispersants and selective reclamation...
Mukhtar et al. CO2 capturing, thermo-kinetic principles, synthesis and amine functionalization of covalent organic polymers for CO2 separation from natural gas: A review
Akeeb et al. Post-combustion CO2 capture via a variety of temperature ranges and material adsorption process: A review
US7947239B2 (en) Carbon dioxide capture and mitigation of carbon dioxide emissions
CA2860684C (en) High purity, high pressure hydrogen production with in-situ co2 and sulfur capture in a single stage reactor
CN104640622B (en) Reproducible absorbent for removing carbon dioxide
Zeng et al. Chemical looping processes—particle characterization, ionic diffusion-reaction mechanism and reactor engineering
US7896953B1 (en) Practical method of CO2 sequestration
CA2543984C (en) Reactivation of lime-based sorbents by co2 shocking
Kotyczka-Moranska et al. Comparison of different methods for enhancing CO2capture by CaO-based sorbents. Review
Tian et al. Simultaneous magnesia regeneration and sulfur dioxide generation in magnesium-based flue gas desulfurization process
Davidson Pre-combustion capture of CO2 in IGCC plants
Dasgupta et al. Robust, high reactivity and enhanced capacity carbon dioxide removal agents for hydrogen production applications
Arstad et al. In-situ XRD studies of dolomite based CO2 sorbents
Iyer High temperature reactive separation process for combined carbon dioxide and sulfur dioxide capture from flue gas and enhanced hydrogen production with in-situ carbon dioxide capture using high reactivity calcium and biomineral sorbents
WO2024080190A1 (en) Carbon dioxide adsorbent, use of carbon dioxide adsorbent, method for isolating carbon dioxide, plant for recovering/reserving carbon dioxide, and method for recovering/reserving carbon dioxide
Iyer et al. High temperature CO2 capture using engineered eggshells: a route to carbon management
Fan et al. Separation of carbon dioxide (CO 2) from gas mixtures
Iyer et al. Enhanced hydrogen production integrated with CO2 separation in a single-stage reactor
Iyer et al. High purity, high pressure hydrogen production with in‐situ CO2 and sulfur capture in a single stage reactor
Trzepizur Waste Ca (Eggshells) natural materials for CO2 capture
Ramkumar et al. Enhanced hydrogen production integrated with CO2 separation in a single-stage reactor

Legal Events

Date Code Title Description
121 Ep: the epo has been informed by wipo that ep was designated in this application
NENP Non-entry into the national phase

Ref country code: DE

122 Ep: pct application non-entry in european phase

Ref document number: 06785920

Country of ref document: EP

Kind code of ref document: A2