Download as pdf or txt
Download as pdf or txt
You are on page 1of 1

Not logged in Talk Contributions Create account Log in

Article Talk Read Edit View history Search Wikipedia

Water-gas shift reaction


From Wikipedia, the free encyclopedia

Main page The water-gas shift reaction (WGSR) describes the reaction of carbon monoxide and water vapor to form carbon dioxide and hydrogen:
Contents
CO + H2O ⇌ CO2 + H2
Current events
Random article 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,
About Wikipedia hydrogen was obtained by reacting steam under high pressure with iron to produce iron, iron oxide and hydrogen. With the development of industrial processes that required hydrogen, such as the Haber–
Contact us
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
Donate
a pure hydrogen product. As the idea of hydrogen economy gains popularity, the focus on hydrogen as a replacement fuel source for hydrocarbons is increasing.
Contribute
Contents [hide]
Help
Community portal 1 Applications
Recent changes 1.1 Fuel cells
Upload file 2 Reaction conditions
2.1 Temperature dependence
Tools
3 Practical concerns
What links here
3.1 Low temperature shift
Related changes
3.2 High temperature shift catalysts
Special pages
Permanent link 4 Mechanism
Page information 4.1 Associative mechanism
Cite this page 4.2 Redox mechanism
Wikidata item 4.3 Homogeneous models

Print/export
4.4 Thermodynamics
5 Reverse water-gas shift
Download as PDF
Printable version 6 See also
7 References
Languages

‫اﻟﻌﺮﺑﻴﺔ‬
Deutsch Applications [ edit ]
Français
The WGSR is an important industrial reaction that is used in the manufacture of ammonia, hydrocarbons, methanol, and hydrogen. It is also often used in conjunction with steam reforming of methane and
Italiano
⽇本語 other hydrocarbons. 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
Português monoxide, which is important for the production of high purity hydrogen for use in ammonia synthesis.
Türkçe
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
4 more water, which suppresses this reaction.
Edit links

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.[1] 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.[1] Unfortunately, current commercial catalysts that are used in
industrial water gas shift processes are not compatible with fuel cell applications.[2] 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.[1] 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.[2] Developing a catalyst that can overcome these limitations is relevant to implementation of a hydrogen economy.

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 carbon monoxide conversion is observed
at lower temperatures.

Temperature dependence [ edit ]

The water gas shift reaction is a moderately exothermic reversible reaction. Therefore, with increasing temperature the reaction rate
increases but the conversion of reactants to products becomes less favorable.[3] Due to its exothermic nature, high carbon monoxide
conversion is thermodynamically favored at low temperatures. Despite the thermodynamic favorability at low temperatures, the reaction
is kinetically favored at high temperatures. The water-gas shift reaction is sensitive to temperature, with a tendency to shift towards
reactants as temperature increases due to Le Chatelier's principle. Over the temperature range 600–2000 K, the equilibrium constant
for the WGSR has the following relationship:[2]

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.[4] The initial HTS takes advantage of the high reaction rates, but is thermodynamically limited, which results in A plot of the temperature dependence of Keq, the value of which
approaches 1 at 1100 K.[2]
incomplete conversion of carbon monoxide and a 2-4% carbon monoxide exit composition. To shift the equilibrium toward hydrogen
production, a subsequent low temperature shift reactor is employed to produce a carbon monoxide exit composition of less than 1%.
The transition from the HTS to the LTS reactors necessitates intersystem cooling. Due to the different reaction conditions, different catalysts must be employed at each stage to ensure optimal activity. The
commercial HTS catalyst is the iron oxide–chromium oxide catalyst and the LTS catalyst is a copper-based catalyst. The order proceeds from high to low temperature due to the susceptibility of the copper
catalyst to poisoning by sulfur that may remain after the steam reformation process.[2] This necessitates the removal of the sulfur compounds prior to the LTS reactor by a guard bed in order to protect the
copper catalyst. Conversely, the iron used in the HTS reaction is generally more robust and resistant toward poisoning by sulfur compounds. While both the HTS and LTS catalysts are commercially
available, their specific composition varies based on vendor. 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.

Low temperature shift [ edit ]

The typical composition of a commercial LTS catalyst has been reported as 32-33% CuO, 34-53% ZnO, 15-33% Al2O3.[2] 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.[5]

High temperature shift catalysts [ 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).[6] The chromium acts to stabilize
the iron oxide and prevents sintering. The operation of HTS catalysts occurs within the temperature range of 310 oC to 450 oC. 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 oC to prevent the exit temperature from exceeding 550 oC. Industrial reactors operate at a range from atmospheric
pressure to 8375 kPa (82.7 atm).[6] 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,[7] with values as low as Ea = 34 kJ/mol reported relative to hydrogen generation.

Mechanism [ edit ]

The WGSR has been extensively studied for over a hundred years. The kinetically relevant mechanism
depends on the catalyst composition and the temperature.[4][11] 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.[3] 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.[12][10]

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
Proposed associative and redox mechanisms of the water gas shift reaction[8][9][10]
mechanism. Experimental studies indicate that both intermediates contribute to the reaction rate over metal
oxide supported transition metal catalysts.[12][10] 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.[10] 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.[13]

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.[1][14]

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.[15]

Reverse water-gas shift [ edit ]

The water-gas shift reaction is an equilibrium reaction, i.e. CO + H2O ⮀ CO2 + H2. 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 the water gas composition (CO:H2 ratio) can be controlled, typically by controlling the steam content. So technically, there is no such thing as ‘reverse’ water-gas shift. It is a
misnomer. Nevertheless, it is widely used to quickly inform an audience of which direction the equilibrium is approached.

See also [ edit ]

In situ resource utilization


Lane hydrogen producer
PROX
Industrial catalysts

References [ edit ]

1. ^ a b c d Vielstich, Wolf; Lamm, Arnold; Gasteiger, Hubert A., eds. 8. ^ Gokhale, Amit A.; Dumesic, James A.; Mavrikakis, Manos (2008- 12. ^ a b Nelson, Nicholas C.; Nguyen, Manh-Thuong; Glezakou,
(2003). Handbook of fuel cells: fundamentals, technology, 01-01). "On the Mechanism of Low-Temperature Water Gas Shift Vassiliki-Alexandra; Rousseau, Roger; Szanyi, János (October
applications. New York: Wiley. ISBN 978-0-471-49926-8. Reaction on Copper". Journal of the American Chemical Society. 2019). "Carboxyl intermediate formation via an in situ-generated

2. ^ a b c d e f Callaghan, Caitlin (2006). Kinetics and catalysis of the 130 (4): 1402–1414. doi:10.1021/ja0768237 . ISSN 0002-7863 . metastable active site during water-gas shift catalysis" . Nature

water-gas-shift reaction: A Microkinetic and Graph Theoretic PMID 18181624 . Catalysis. 2 (10): 916–924. doi:10.1038/s41929-019-0343-2 .

Approach (PDF) (PhD). Worcester Polytechnic Institute. 9. ^ Grabow, Lars C.; Gokhale, Amit A.; Evans, Steven T.; Dumesic, ISSN 2520-1158 . S2CID 202729116 .

3. ^ a b Ratnasamy, Chandra; Wagner, Jon P. (September 2009). James A.; Mavrikakis, Manos (2008-03-01). "Mechanism of the 13. ^ Nelson, Nicholas C.; Chen, Linxiao; Meira, Debora; Kovarik,
"Water Gas Shift Catalysis". Catalysis Reviews. 51 (3): 325–440. Water Gas Shift Reaction on Pt: First Principles, Experiments, and Libor; Szanyi, János (2020). "In Situ Dispersion of Palladium on
doi:10.1080/01614940903048661 . S2CID 98530242 . Microkinetic Modeling". The Journal of Physical Chemistry C. 112 TiO2 During Reverse Water–Gas Shift Reaction: Formation of
(12): 4608–4617. doi:10.1021/jp7099702 . ISSN 1932-7447 . Atomically Dispersed Palladium". Angewandte Chemie
4. ^ a b Smith R J, Byron; Muruganandam Loganthan; Murthy
Shekhar Shantha (2010). "A Review of the Water Gas Shift 10. ^ a b c d Nelson, Nicholas C.; Szanyi, János (2020-05-15). International Edition. n/a (n/a). doi:10.1002/anie.202007576 .
"Heterolytic Hydrogen Activation: Understanding Support Effects in ISSN 1521-3773 . PMID 32589820 .
Reaction". International Journal of Chemical Reactor Engineering.
8: 1–32. doi:10.2202/1542-6580.2238 . S2CID 96769998 . Water–Gas Shift, Hydrodeoxygenation, and CO Oxidation 14. ^ Barakat, Tarek; Rooke, Joanna C.; Genty, Eric; Cousin, Renaud;
Catalysis". ACS Catalysis. 10 (10): 5663–5671. Siffert, Stéphane; Su, Bao-Lian (1 January 2013). "Gold catalysts
5. ^ Jain, Rishabh; Maric, Radenka (April 2014). "Synthesis of nano-
doi:10.1021/acscatal.0c01059 . in environmental remediation and water-gas shift technologies".
Pt onto ceria support as catalyst for water–gas shift reaction by
11. ^ Yao, Siyu; Zhang, Xiao; Zhou, Wu; Gao, Rui; Xu, Wenqian; Ye, Energy & Environmental Science. 6 (2): 371.
Reactive Spray Deposition Technology". Applied Catalysis A:
Yifan; Lin, Lili; Wen, Xiaodong; Liu, Ping; Chen, Bingbing; Crumlin, doi:10.1039/c2ee22859a .
General. 475: 461–468. doi:10.1016/j.apcata.2014.01.053 .
Ethan (2017-06-22). "Atomic-layered Au clusters on α-MoC as 15. ^ King, A. D.; King, R. B.; Yang, D. B., "Homogeneous catalysis of
6. ^ a b Newsome, David S. (1980). "The Water-Gas Shift Reaction".
catalysts for the low-temperature water-gas shift reaction" . the water gas shift reaction using iron pentacarbonyl", J. Am.
Catalysis Reviews: Science and Engineering. 21 (2): 275–318.
Science. 357 (6349): 389–393. doi:10.1126/science.aah4321 . Chem. Soc. 1980, vol. 102, pp. 1028-1032.
doi:10.1080/03602458008067535 .
ISSN 0036-8075 . PMID 28642235 . S2CID 206651887 . doi:10.1021/ja00523a020
7. ^ 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 .

Categories: Inorganic reactions Chemical processes Hydrogen production Industrial gases

This page was last edited on 16 September 2020, at 13:49 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia
Foundation, Inc., a non-profit organization.

Privacy policy About Wikipedia Disclaimers Contact Wikipedia Mobile view Developers Statistics Cookie statement

You might also like