fbpx
Wikipedia

Water–gas shift reaction

The water–gas shift reaction (WGSR) describes the reaction of carbon monoxide and water vapor to form carbon dioxide and hydrogen:

CO + H2O ⇌ CO2 + H2

The water gas shift reaction was discovered by Italian physicist Felice Fontana in 1780. It was not until much later that the industrial value of this reaction was realized. Before the early 20th century, hydrogen was obtained by reacting steam under high pressure with iron to produce iron oxide and hydrogen. With the development of industrial processes that required hydrogen, such as the Haber–Bosch ammonia synthesis, a less expensive and more efficient method of hydrogen production was needed. As a resolution to this problem, the WGSR was combined with the gasification of coal to produce hydrogen.

Applications edit

The WGSR is a highly valuable industrial reaction that is used in the manufacture of ammonia, hydrocarbons, methanol, and hydrogen. Its most important application is in conjunction with the conversion of carbon monoxide from steam reforming of methane or other hydrocarbons in the production of hydrogen.[1] In the Fischer–Tropsch process, the WGSR is one of the most important reactions used to balance the H2/CO ratio. It provides a source of hydrogen at the expense of carbon monoxide, which is important for the production of high purity hydrogen for use in ammonia synthesis.

The water–gas shift reaction may be an undesired side reaction in processes involving water and carbon monoxide, e.g. the rhodium-based Monsanto process. The iridium-based Cativa process uses less water, which suppresses this reaction.

Fuel cells edit

The WGSR can aid in the efficiency of fuel cells by increasing hydrogen production. The WGSR is considered a critical component in the reduction of carbon monoxide concentrations in cells that are susceptible to carbon monoxide poisoning such as the proton-exchange membrane (PEM) fuel cell.[2] The benefits of this application are two-fold: not only would the water gas shift reaction effectively reduce the concentration of carbon monoxide, but it would also increase the efficiency of the fuel cells by increasing hydrogen production.[2] Unfortunately, current commercial catalysts that are used in industrial water gas shift processes are not compatible with fuel cell applications.[3] With the high demand for clean fuel and the critical role of the water gas shift reaction in hydrogen fuel cells, the development of water gas shift catalysts for the application in fuel cell technology is an area of current research interest.

Catalysts for fuel cell application would need to operate at low temperatures. Since the WGSR is slow at lower temperatures where equilibrium favors hydrogen production, WGS reactors require large amounts of catalysts, which increases their cost and size beyond practical application.[2] The commercial LTS catalyst used in large scale industrial plants is also pyrophoric in its inactive state and therefore presents safety concerns for consumer applications.[3] Developing a catalyst that can overcome these limitations is relevant to implementation of a hydrogen economy.

Sorption enhanced water gas shift edit

The WGS reaction is used in combination with the solid adsorption of CO2 in the sorption enhanced water gas shift (SEWGS) in order to produce a high pressure hydrogen stream from syngas.[4]

Reaction conditions edit

The equilibrium of this reaction shows a significant temperature dependence and the equilibrium constant decreases with an increase in temperature, that is, higher hydrogen formation is observed at lower temperatures.

Temperature dependence edit

 
Temperature dependence of the free molar (Gibbs) enthalpy and equilibrium constant of the water-gas shift reaction.

With increasing temperature, the reaction rate increases, but hydrogen production becomes less favorable thermodynamically[5] since the water gas shift reaction is moderately exothermic; this shift in chemical equilibrium can be explained according to Le Chatelier's principle. Over the temperature range of 600–2000 K, the equilibrium constant for the WGSR has the following relationship:[3]

 

Practical concerns edit

In order to take advantage of both the thermodynamics and kinetics of the reaction, the industrial scale water gas shift reaction is conducted in multiple adiabatic stages consisting of a high temperature shift (HTS) followed by a low temperature shift (LTS) with intersystem cooling.[6] The initial HTS takes advantage of the high reaction rates, but results in incomplete conversion of carbon monoxide. A subsequent low temperature shift reactor lowers the carbon monoxide content to <1%. Commercial HTS catalysts are based on iron oxidechromium oxide and the LTS catalyst is a copper-based. The copper catalyst is susceptible to poisoning by sulfur. Sulfur compounds are removed prior to the LTS reactor by a guard bed. An important limitation for the HTS is the H2O/CO ratio where low ratios may lead to side reactions such as the formation of metallic iron, methanation, carbon deposition, and the Fischer–Tropsch reaction.

High temperature shift catalysis edit

The typical composition of commercial HTS catalyst has been reported as 74.2% Fe2O3, 10.0% Cr2O3, 0.2% MgO (remaining percentage attributed to volatile components).[7] The chromium acts to stabilize the iron oxide and prevents sintering. The operation of HTS catalysts occurs within the temperature range of 310 °C to 450 °C. The temperature increases along the length of the reactor due to the exothermic nature of the reaction. As such, the inlet temperature is maintained at 350 °C to prevent the exit temperature from exceeding 550 °C. Industrial reactors operate at a range from atmospheric pressure to 8375 kPa (82.7 atm).[7] The search for high performance HT WGS catalysts remains an intensive topic of research in fields of chemistry and materials science. Activation energy is a key criteria for the assessment of catalytic performance in WGS reactions. To date, some of the lowest activation energy values have been found for catalysts consisting of copper nanoparticles on ceria support materials,[8] with values as low as Ea = 34 kJ/mol reported relative to hydrogen generation.

Low temperature shift catalysis edit

Catalysts for the lower temperature WGS reaction are commonly based on copper or copper oxide loaded ceramic phases, While the most common supports include alumina or alumina with zinc oxide, other supports may include rare earth oxides, spinels or perovskites.[9] A typical composition of a commercial LTS catalyst has been reported as 32-33% CuO, 34-53% ZnO, 15-33% Al2O3.[3] The active catalytic species is CuO. The function of ZnO is to provide structural support as well as prevent the poisoning of copper by sulfur. The Al2O3 prevents dispersion and pellet shrinkage. The LTS shift reactor operates at a range of 200–250 °C. The upper temperature limit is due to the susceptibility of copper to thermal sintering. These lower temperatures also reduce the occurrence of side reactions that are observed in the case of the HTS. Noble metals such as platinum, supported on ceria, have also been used for LTS.[10]

Mechanism edit

 
Proposed associative and redox mechanisms of the water gas shift reaction[11][12][13]

The WGSR has been extensively studied for over a hundred years. The kinetically relevant mechanism depends on the catalyst composition and the temperature.[6][14] Two mechanisms have been proposed: an associative Langmuir–Hinshelwood mechanism and a redox mechanism. The redox mechanism is generally regarded as kinetically relevant during the high-temperature WGSR (> 350 °C) over the industrial iron-chromia catalyst.[5] Historically, there has been much more controversy surrounding the mechanism at low temperatures. Recent experimental studies confirm that the associative carboxyl mechanism is the predominant low temperature pathway on metal-oxide-supported transition metal catalysts.[15][13]

Associative mechanism edit

In 1920 Armstrong and Hilditch first proposed the associative mechanism. In this mechanism CO and H2O are adsorbed onto the surface of the catalyst, followed by formation of an intermediate and the desorption of H2 and CO2. In general, H2O dissociates onto the catalyst to yield adsorbed OH and H. The dissociated water reacts with CO to form a carboxyl or formate intermediate. The intermediate subsequently dehydrogenates to yield CO2 and adsorbed H. Two adsorbed H atoms recombine to form H2.

There has been significant controversy surrounding the kinetically relevant intermediate during the associative mechanism. Experimental studies indicate that both intermediates contribute to the reaction rate over metal oxide supported transition metal catalysts.[15][13] However, the carboxyl pathway accounts for about 90% of the total rate owing to the thermodynamic stability of adsorbed formate on the oxide support. The active site for carboxyl formation consists of a metal atom adjacent to an adsorbed hydroxyl. This ensemble is readily formed at the metal-oxide interface and explains the much higher activity of oxide-supported transition metals relative to extended metal surfaces.[13] The turn-over-frequency for the WGSR is proportional to the equilibrium constant of hydroxyl formation, which rationalizes why reducible oxide supports (e.g. CeO2) are more active than irreducible supports (e.g. SiO2) and extended metal surfaces (e.g. Pt). In contrast to the active site for carboxyl formation, formate formation occurs on extended metal surfaces. The formate intermediate can be eliminated during the WGSR by using oxide-supported atomically dispersed transition metal catalysts, further confirming the kinetic dominance of the carboxyl pathway.[16]

Redox mechanism edit

The redox mechanism involves a change in the oxidation state of the catalytic material. In this mechanism, CO is oxidized by an O-atom intrinsically belonging to the catalytic material to form CO2. A water molecule undergoes dissociative adsorption at the newly formed O-vacancy to yield two hydroxyls. The hydroxyls disproportionate to yield H2 and return the catalytic surface back to its pre-reaction state.

Homogeneous models edit

The mechanism entails nucleophilic attack of water or hydroxide on a M-CO center, generating a metallacarboxylic acid.[2][17]

Thermodynamics edit

The WGSR is exergonic, with the following thermodynamic parameters at room temperature (298 K):

Free energy ΔG = –6.82 kcal
Enthalpy ΔH = –9.84 kcal
Entropy ΔS = –10.1 cal/deg

In aqueous solution, the reaction is less exergonic.[18]

Reverse water–gas shift edit

In the conversion of carbon dioxide to useful materials, the water–gas shift reaction is used to produce carbon monoxide from hydrogen and carbon dioxide. This is sometimes called the reverse water–gas shift reaction.[19]

Water gas is defined as a fuel gas consisting mainly of carbon monoxide (CO) and hydrogen (H2). The term ‘shift’ in water–gas shift means changing the water gas composition (CO:H2) ratio. The ratio can be increased by adding CO2 or reduced by adding steam to the reactor.

See also edit

References edit

  1. ^ Water Gas Shift Catalysis a combined experimental and computational study
  2. ^ a b c d Vielstich, Wolf; Lamm, Arnold; Gasteiger, Hubert A., eds. (2003). Handbook of fuel cells: fundamentals, technology, applications. New York: Wiley. ISBN 978-0-471-49926-8.
  3. ^ a b c d Callaghan, Caitlin (2006). Kinetics and catalysis of the water-gas-shift reaction: A Microkinetic and Graph Theoretic Approach (PDF) (PhD). Worcester Polytechnic Institute.
  4. ^ Jansen, Daniel; van Selow, Edward; Cobden, Paul; Manzolini, Giampaolo; Macchi, Ennio; Gazzani, Matteo; Blom, Richard; Heriksen, Partow Pakdel; Beavis, Rich; Wright, Andrew (2013-01-01). "SEWGS Technology is Now Ready for Scale-up!" (PDF). Energy Procedia. 37: 2265–2273. doi:10.1016/j.egypro.2013.06.107. ISSN 1876-6102.
  5. ^ a b Ratnasamy, Chandra; Wagner, Jon P. (September 2009). "Water Gas Shift Catalysis". Catalysis Reviews. 51 (3): 325–440. doi:10.1080/01614940903048661. S2CID 98530242.
  6. ^ a b Smith R J, Byron; Muruganandam Loganthan; Murthy Shekhar Shantha (2010). "A Review of the Water Gas Shift Reaction". International Journal of Chemical Reactor Engineering. 8: 1–32. doi:10.2202/1542-6580.2238. S2CID 96769998.
  7. ^ a b Newsome, David S. (1980). "The Water-Gas Shift Reaction". Catalysis Reviews: Science and Engineering. 21 (2): 275–318. doi:10.1080/03602458008067535.
  8. ^ Rodriguez, J.A.; Liu, P.; Wang, X.; Wen, W.; Hanson, J.; Hrbek, J.; Pérez, M.; Evans, J. (15 May 2009). "Water-gas shift activity of Cu surfaces and Cu nanoparticles supported on metal oxides". Catalysis Today. 143 (1–2): 45–50. doi:10.1016/j.cattod.2008.08.022.
  9. ^ Coletta, Vitor C.; Gonçalves, Renato V.; Bernardi, Maria I. B.; Hanaor, Dorian A. H.; Assadi, M. Hussein N.; Marcos, Francielle C. F.; Nogueira, Francisco G. E.; Assaf, Elisabete M.; Mastelaro, Valmor R. (2021). "Cu-Modified SrTiO3 Perovskites Toward Enhanced Water–Gas Shift Catalysis: A Combined Experimental and Computational Study". ACS Applied Energy Materials. 4: 452–461. arXiv:2104.06739. doi:10.1021/acsaem.0c02371. S2CID 233231670.
  10. ^ Jain, Rishabh; Maric, Radenka (April 2014). "Synthesis of nano-Pt onto ceria support as catalyst for water–gas shift reaction by Reactive Spray Deposition Technology". Applied Catalysis A: General. 475: 461–468. doi:10.1016/j.apcata.2014.01.053.
  11. ^ Gokhale, Amit A.; Dumesic, James A.; Mavrikakis, Manos (2008-01-01). "On the Mechanism of Low-Temperature Water Gas Shift Reaction on Copper". Journal of the American Chemical Society. 130 (4): 1402–1414. doi:10.1021/ja0768237. ISSN 0002-7863. PMID 18181624.
  12. ^ Grabow, Lars C.; Gokhale, Amit A.; Evans, Steven T.; Dumesic, James A.; Mavrikakis, Manos (2008-03-01). "Mechanism of the Water Gas Shift Reaction on Pt: First Principles, Experiments, and Microkinetic Modeling". The Journal of Physical Chemistry C. 112 (12): 4608–4617. doi:10.1021/jp7099702. ISSN 1932-7447.
  13. ^ a b c d Nelson, Nicholas C.; Szanyi, János (2020-05-15). "Heterolytic Hydrogen Activation: Understanding Support Effects in Water–Gas Shift, Hydrodeoxygenation, and CO Oxidation Catalysis". ACS Catalysis. 10 (10): 5663–5671. doi:10.1021/acscatal.0c01059. OSTI 1656557. S2CID 218798723.
  14. ^ Yao, Siyu; Zhang, Xiao; Zhou, Wu; Gao, Rui; Xu, Wenqian; Ye, Yifan; Lin, Lili; Wen, Xiaodong; Liu, Ping; Chen, Bingbing; Crumlin, Ethan (2017-06-22). "Atomic-layered Au clusters on α-MoC as catalysts for the low-temperature water-gas shift reaction" (PDF). Science. 357 (6349): 389–393. Bibcode:2017Sci...357..389Y. doi:10.1126/science.aah4321. ISSN 0036-8075. PMID 28642235. S2CID 206651887.
  15. ^ a b Nelson, Nicholas C.; Nguyen, Manh-Thuong; Glezakou, Vassiliki-Alexandra; Rousseau, Roger; Szanyi, János (October 2019). "Carboxyl intermediate formation via an in situ-generated metastable active site during water-gas shift catalysis". Nature Catalysis. 2 (10): 916–924. doi:10.1038/s41929-019-0343-2. ISSN 2520-1158. S2CID 202729116.
  16. ^ Nelson, Nicholas C.; Chen, Linxiao; Meira, Debora; Kovarik, Libor; Szanyi, János (2020). "In Situ Dispersion of Palladium on TiO2 During Reverse Water–Gas Shift Reaction: Formation of Atomically Dispersed Palladium". Angewandte Chemie International Edition. 59 (40): 17657–17663. doi:10.1002/anie.202007576. ISSN 1521-3773. OSTI 1661896. PMID 32589820. S2CID 220118889.
  17. ^ Barakat, Tarek; Rooke, Joanna C.; Genty, Eric; Cousin, Renaud; Siffert, Stéphane; Su, Bao-Lian (1 January 2013). "Gold catalysts in environmental remediation and water-gas shift technologies". Energy & Environmental Science. 6 (2): 371. doi:10.1039/c2ee22859a.
  18. ^ King, A. D.; King, R. B.; Yang, D. B., "Homogeneous catalysis of the water gas shift reaction using iron pentacarbonyl", J. Am. Chem. Soc. 1980, vol. 102, pp. 1028-1032. doi:10.1021/ja00523a020
  19. ^ Guil-López, R.; Mota, N.; Llorente, J.; Millán, E.; Pawelec, B.; Fierro, J. L. G.; Navarro, R. M. (2019). "Methanol Synthesis from CO2: A Review of the Latest Developments in Heterogeneous Catalysis". Materials. 12 (23): 3902. Bibcode:2019Mate...12.3902G. doi:10.3390/ma12233902. ISSN 1996-1944. PMC 6926878. PMID 31779127.

water, shift, reaction, water, shift, reaction, wgsr, describes, reaction, carbon, monoxide, water, vapor, form, carbon, dioxide, hydrogen, water, shift, reaction, discovered, italian, physicist, felice, fontana, 1780, until, much, later, that, industrial, val. The water gas shift reaction WGSR describes the reaction of carbon monoxide and water vapor to form carbon dioxide and hydrogen CO H2O CO2 H2 The water gas shift reaction was discovered by Italian physicist Felice Fontana in 1780 It was not until much later that the industrial value of this reaction was realized Before the early 20th century hydrogen was obtained by reacting steam under high pressure with iron to produce iron oxide and hydrogen With the development of industrial processes that required hydrogen such as the Haber Bosch ammonia synthesis a less expensive and more efficient method of hydrogen production was needed As a resolution to this problem the WGSR was combined with the gasification of coal to produce hydrogen Contents 1 Applications 1 1 Fuel cells 1 2 Sorption enhanced water gas shift 2 Reaction conditions 2 1 Temperature dependence 3 Practical concerns 3 1 High temperature shift catalysis 3 2 Low temperature shift catalysis 4 Mechanism 4 1 Associative mechanism 4 2 Redox mechanism 4 3 Homogeneous models 4 4 Thermodynamics 5 Reverse water gas shift 6 See also 7 ReferencesApplications editThe WGSR is a highly valuable industrial reaction that is used in the manufacture of ammonia hydrocarbons methanol and hydrogen Its most important application is in conjunction with the conversion of carbon monoxide from steam reforming of methane or other hydrocarbons in the production of hydrogen 1 In the Fischer Tropsch process the WGSR is one of the most important reactions used to balance the H2 CO ratio It provides a source of hydrogen at the expense of carbon monoxide which is important for the production of high purity hydrogen for use in ammonia synthesis The water gas shift reaction may be an undesired side reaction in processes involving water and carbon monoxide e g the rhodium based Monsanto process The iridium based Cativa process uses less water which suppresses this reaction Fuel cells edit The WGSR can aid in the efficiency of fuel cells by increasing hydrogen production The WGSR is considered a critical component in the reduction of carbon monoxide concentrations in cells that are susceptible to carbon monoxide poisoning such as the proton exchange membrane PEM fuel cell 2 The benefits of this application are two fold not only would the water gas shift reaction effectively reduce the concentration of carbon monoxide but it would also increase the efficiency of the fuel cells by increasing hydrogen production 2 Unfortunately current commercial catalysts that are used in industrial water gas shift processes are not compatible with fuel cell applications 3 With the high demand for clean fuel and the critical role of the water gas shift reaction in hydrogen fuel cells the development of water gas shift catalysts for the application in fuel cell technology is an area of current research interest Catalysts for fuel cell application would need to operate at low temperatures Since the WGSR is slow at lower temperatures where equilibrium favors hydrogen production WGS reactors require large amounts of catalysts which increases their cost and size beyond practical application 2 The commercial LTS catalyst used in large scale industrial plants is also pyrophoric in its inactive state and therefore presents safety concerns for consumer applications 3 Developing a catalyst that can overcome these limitations is relevant to implementation of a hydrogen economy Sorption enhanced water gas shift edit The WGS reaction is used in combination with the solid adsorption of CO2 in the sorption enhanced water gas shift SEWGS in order to produce a high pressure hydrogen stream from syngas 4 Reaction conditions editThe equilibrium of this reaction shows a significant temperature dependence and the equilibrium constant decreases with an increase in temperature that is higher hydrogen formation is observed at lower temperatures Temperature dependence edit nbsp Temperature dependence of the free molar Gibbs enthalpy and equilibrium constant of the water gas shift reaction With increasing temperature the reaction rate increases but hydrogen production becomes less favorable thermodynamically 5 since the water gas shift reaction is moderately exothermic this shift in chemical equilibrium can be explained according to Le Chatelier s principle Over the temperature range of 600 2000 K the equilibrium constant for the WGSR has the following relationship 3 K e q 10 2 4198 0 0003855 T 2180 6 T displaystyle K mathrm eq 10 2 4198 0 0003855T frac 2180 6 T nbsp Practical concerns editIn order to take advantage of both the thermodynamics and kinetics of the reaction the industrial scale water gas shift reaction is conducted in multiple adiabatic stages consisting of a high temperature shift HTS followed by a low temperature shift LTS with intersystem cooling 6 The initial HTS takes advantage of the high reaction rates but results in incomplete conversion of carbon monoxide A subsequent low temperature shift reactor lowers the carbon monoxide content to lt 1 Commercial HTS catalysts are based on iron oxide chromium oxide and the LTS catalyst is a copper based The copper catalyst is susceptible to poisoning by sulfur Sulfur compounds are removed prior to the LTS reactor by a guard bed An important limitation for the HTS is the H2O CO ratio where low ratios may lead to side reactions such as the formation of metallic iron methanation carbon deposition and the Fischer Tropsch reaction High temperature shift catalysis edit The typical composition of commercial HTS catalyst has been reported as 74 2 Fe2O3 10 0 Cr2O3 0 2 MgO remaining percentage attributed to volatile components 7 The chromium acts to stabilize the iron oxide and prevents sintering The operation of HTS catalysts occurs within the temperature range of 310 C to 450 C The temperature increases along the length of the reactor due to the exothermic nature of the reaction As such the inlet temperature is maintained at 350 C to prevent the exit temperature from exceeding 550 C Industrial reactors operate at a range from atmospheric pressure to 8375 kPa 82 7 atm 7 The search for high performance HT WGS catalysts remains an intensive topic of research in fields of chemistry and materials science Activation energy is a key criteria for the assessment of catalytic performance in WGS reactions To date some of the lowest activation energy values have been found for catalysts consisting of copper nanoparticles on ceria support materials 8 with values as low as Ea 34 kJ mol reported relative to hydrogen generation Low temperature shift catalysis edit Catalysts for the lower temperature WGS reaction are commonly based on copper or copper oxide loaded ceramic phases While the most common supports include alumina or alumina with zinc oxide other supports may include rare earth oxides spinels or perovskites 9 A typical composition of a commercial LTS catalyst has been reported as 32 33 CuO 34 53 ZnO 15 33 Al2O3 3 The active catalytic species is CuO The function of ZnO is to provide structural support as well as prevent the poisoning of copper by sulfur The Al2O3 prevents dispersion and pellet shrinkage The LTS shift reactor operates at a range of 200 250 C The upper temperature limit is due to the susceptibility of copper to thermal sintering These lower temperatures also reduce the occurrence of side reactions that are observed in the case of the HTS Noble metals such as platinum supported on ceria have also been used for LTS 10 Mechanism edit nbsp Proposed associative and redox mechanisms of the water gas shift reaction 11 12 13 The WGSR has been extensively studied for over a hundred years The kinetically relevant mechanism depends on the catalyst composition and the temperature 6 14 Two mechanisms have been proposed an associative Langmuir Hinshelwood mechanism and a redox mechanism The redox mechanism is generally regarded as kinetically relevant during the high temperature WGSR gt 350 C over the industrial iron chromia catalyst 5 Historically there has been much more controversy surrounding the mechanism at low temperatures Recent experimental studies confirm that the associative carboxyl mechanism is the predominant low temperature pathway on metal oxide supported transition metal catalysts 15 13 Associative mechanism edit In 1920 Armstrong and Hilditch first proposed the associative mechanism In this mechanism CO and H2O are adsorbed onto the surface of the catalyst followed by formation of an intermediate and the desorption of H2 and CO2 In general H2O dissociates onto the catalyst to yield adsorbed OH and H The dissociated water reacts with CO to form a carboxyl or formate intermediate The intermediate subsequently dehydrogenates to yield CO2 and adsorbed H Two adsorbed H atoms recombine to form H2 There has been significant controversy surrounding the kinetically relevant intermediate during the associative mechanism Experimental studies indicate that both intermediates contribute to the reaction rate over metal oxide supported transition metal catalysts 15 13 However the carboxyl pathway accounts for about 90 of the total rate owing to the thermodynamic stability of adsorbed formate on the oxide support The active site for carboxyl formation consists of a metal atom adjacent to an adsorbed hydroxyl This ensemble is readily formed at the metal oxide interface and explains the much higher activity of oxide supported transition metals relative to extended metal surfaces 13 The turn over frequency for the WGSR is proportional to the equilibrium constant of hydroxyl formation which rationalizes why reducible oxide supports e g CeO2 are more active than irreducible supports e g SiO2 and extended metal surfaces e g Pt In contrast to the active site for carboxyl formation formate formation occurs on extended metal surfaces The formate intermediate can be eliminated during the WGSR by using oxide supported atomically dispersed transition metal catalysts further confirming the kinetic dominance of the carboxyl pathway 16 Redox mechanism edit The redox mechanism involves a change in the oxidation state of the catalytic material In this mechanism CO is oxidized by an O atom intrinsically belonging to the catalytic material to form CO2 A water molecule undergoes dissociative adsorption at the newly formed O vacancy to yield two hydroxyls The hydroxyls disproportionate to yield H2 and return the catalytic surface back to its pre reaction state Homogeneous models edit The mechanism entails nucleophilic attack of water or hydroxide on a M CO center generating a metallacarboxylic acid 2 17 Thermodynamics edit The WGSR is exergonic with the following thermodynamic parameters at room temperature 298 K Free energy DG 6 82 kcal Enthalpy DH 9 84 kcal Entropy DS 10 1 cal deg In aqueous solution the reaction is less exergonic 18 Reverse water gas shift editIn the conversion of carbon dioxide to useful materials the water gas shift reaction is used to produce carbon monoxide from hydrogen and carbon dioxide This is sometimes called the reverse water gas shift reaction 19 Water gas is defined as a fuel gas consisting mainly of carbon monoxide CO and hydrogen H2 The term shift in water gas shift means changing the water gas composition CO H2 ratio The ratio can be increased by adding CO2 or reduced by adding steam to the reactor See also editIn situ resource utilization Lane hydrogen producer PROX Industrial catalysts Sorption enhanced water gas shift SyngasReferences edit Water Gas Shift Catalysis a combined experimental and computational study a b c d Vielstich Wolf Lamm Arnold Gasteiger Hubert A eds 2003 Handbook of fuel cells fundamentals technology applications New York Wiley ISBN 978 0 471 49926 8 a b c d Callaghan Caitlin 2006 Kinetics and catalysis of the water gas shift reaction A Microkinetic and Graph Theoretic Approach PDF PhD Worcester Polytechnic Institute Jansen Daniel van Selow Edward Cobden Paul Manzolini Giampaolo Macchi Ennio Gazzani Matteo Blom Richard Heriksen Partow Pakdel Beavis Rich Wright Andrew 2013 01 01 SEWGS Technology is Now Ready for Scale up PDF Energy Procedia 37 2265 2273 doi 10 1016 j egypro 2013 06 107 ISSN 1876 6102 a b Ratnasamy Chandra Wagner Jon P September 2009 Water Gas Shift Catalysis Catalysis Reviews 51 3 325 440 doi 10 1080 01614940903048661 S2CID 98530242 a b Smith R J Byron Muruganandam Loganthan Murthy Shekhar Shantha 2010 A Review of the Water Gas Shift Reaction International Journal of Chemical Reactor Engineering 8 1 32 doi 10 2202 1542 6580 2238 S2CID 96769998 a b Newsome David S 1980 The Water Gas Shift Reaction Catalysis Reviews Science and Engineering 21 2 275 318 doi 10 1080 03602458008067535 Rodriguez J A Liu P Wang X Wen W Hanson J Hrbek J Perez M Evans J 15 May 2009 Water gas shift activity of Cu surfaces and Cu nanoparticles supported on metal oxides Catalysis Today 143 1 2 45 50 doi 10 1016 j cattod 2008 08 022 Coletta Vitor C Goncalves Renato V Bernardi Maria I B Hanaor Dorian A H Assadi M Hussein N Marcos Francielle C F Nogueira Francisco G E Assaf Elisabete M Mastelaro Valmor R 2021 Cu Modified SrTiO3 Perovskites Toward Enhanced Water Gas Shift Catalysis A Combined Experimental and Computational Study ACS Applied Energy Materials 4 452 461 arXiv 2104 06739 doi 10 1021 acsaem 0c02371 S2CID 233231670 Jain Rishabh Maric Radenka April 2014 Synthesis of nano Pt onto ceria support as catalyst for water gas shift reaction by Reactive Spray Deposition Technology Applied Catalysis A General 475 461 468 doi 10 1016 j apcata 2014 01 053 Gokhale Amit A Dumesic James A Mavrikakis Manos 2008 01 01 On the Mechanism of Low Temperature Water Gas Shift Reaction on Copper Journal of the American Chemical Society 130 4 1402 1414 doi 10 1021 ja0768237 ISSN 0002 7863 PMID 18181624 Grabow Lars C Gokhale Amit A Evans Steven T Dumesic James A Mavrikakis Manos 2008 03 01 Mechanism of the Water Gas Shift Reaction on Pt First Principles Experiments and Microkinetic Modeling The Journal of Physical Chemistry C 112 12 4608 4617 doi 10 1021 jp7099702 ISSN 1932 7447 a b c d Nelson Nicholas C Szanyi Janos 2020 05 15 Heterolytic Hydrogen Activation Understanding Support Effects in Water Gas Shift Hydrodeoxygenation and CO Oxidation Catalysis ACS Catalysis 10 10 5663 5671 doi 10 1021 acscatal 0c01059 OSTI 1656557 S2CID 218798723 Yao Siyu Zhang Xiao Zhou Wu Gao Rui Xu Wenqian Ye Yifan Lin Lili Wen Xiaodong Liu Ping Chen Bingbing Crumlin Ethan 2017 06 22 Atomic layered Au clusters on a MoC as catalysts for the low temperature water gas shift reaction PDF Science 357 6349 389 393 Bibcode 2017Sci 357 389Y doi 10 1126 science aah4321 ISSN 0036 8075 PMID 28642235 S2CID 206651887 a b Nelson Nicholas C Nguyen Manh Thuong Glezakou Vassiliki Alexandra Rousseau Roger Szanyi Janos October 2019 Carboxyl intermediate formation via an in situ generated metastable active site during water gas shift catalysis Nature Catalysis 2 10 916 924 doi 10 1038 s41929 019 0343 2 ISSN 2520 1158 S2CID 202729116 Nelson Nicholas C Chen Linxiao Meira Debora Kovarik Libor Szanyi Janos 2020 In Situ Dispersion of Palladium on TiO2 During Reverse Water Gas Shift Reaction Formation of Atomically Dispersed Palladium Angewandte Chemie International Edition 59 40 17657 17663 doi 10 1002 anie 202007576 ISSN 1521 3773 OSTI 1661896 PMID 32589820 S2CID 220118889 Barakat Tarek Rooke Joanna C Genty Eric Cousin Renaud Siffert Stephane Su Bao Lian 1 January 2013 Gold catalysts in environmental remediation and water gas shift technologies Energy amp Environmental Science 6 2 371 doi 10 1039 c2ee22859a King A D King R B Yang D B Homogeneous catalysis of the water gas shift reaction using iron pentacarbonyl J Am Chem Soc 1980 vol 102 pp 1028 1032 doi 10 1021 ja00523a020 Guil Lopez R Mota N Llorente J Millan E Pawelec B Fierro J L G Navarro R M 2019 Methanol Synthesis from CO2 A Review of the Latest Developments in Heterogeneous Catalysis Materials 12 23 3902 Bibcode 2019Mate 12 3902G doi 10 3390 ma12233902 ISSN 1996 1944 PMC 6926878 PMID 31779127 Retrieved from https en wikipedia org w index php title Water gas shift reaction amp oldid 1213580385 Catalysts, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.