Next Article in Journal
Recent Advances in Cobalt and Related Catalysts: From Catalyst Preparation to Catalytic Performance
Previous Article in Journal
Low-Dimensional Nanostructured Photocatalysts for Efficient CO2 Conversion into Solar Fuels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application Prospect of K Used for Catalytic Removal of NOx, COx, and VOCs from Industrial Flue Gas: A Review

1
CAS Key Laboratory of Green Process and Engineering, Institute of Process Engineering, Innovation Academy for Green Manufacture, Chinese Academy of Sciences, Beijing 100190, China
2
School of Chemical Engineering, University of Chinese Academy of Sciences, Beijing 100049, China
3
Beijing Municipal Research Institute of Environmental Protection, Water and Ecology Institution, 59th, Beiyingfang Mid. Street, Xicheng District, Beijing 100037, China
4
College of Environmental Science and Engineering, Beijing Forestry University, 35 Qinghua East Road, Haidian District, Beijing 100083, China
5
Center for Excellence in Regional Atmospheric Environment, Institute of Urban Environment, Chinese Academy of Sciences, Xiamen 361021, China
*
Authors to whom correspondence should be addressed.
Catalysts 2021, 11(4), 419; https://doi.org/10.3390/catal11040419
Submission received: 1 March 2021 / Revised: 22 March 2021 / Accepted: 23 March 2021 / Published: 25 March 2021
(This article belongs to the Section Environmental Catalysis)

Abstract

:
NOx, COx, and volatile organic compounds (VOCs) widely exist in motor vehicle exhaust, coke oven flue gas, sintering flue gas, and pelletizing flue gas. Potassium species have an excellent promotion effect on various catalytic reactions for the treatment of these pollutants. This work reviews the promotion effects of potassium species on the reaction processes, including adsorption, desorption, the pathway and selectivity of reaction, recovery of active center, and effects on the properties of catalysts, including basicity, electron donor characteristics, redox property, active center, stability, and strong metal-to support interaction. The suggestions about how to improve the promotion effects of potassium species in various catalytic reactions are put forward, which involve controlling carriers, content, preparation methods and reaction conditions. The promotion effects of different alkali metals are also compared. The article number about commonly used active metals and promotion ways are also analyzed by bibliometric on NOx, COx, and VOCs. The promotion mechanism of potassium species on various reactions is similar; therefore, the application prospect of potassium species for the coupling control of multi-pollutants in industrial flue gas at low-temperature is described.

1. Introduction

NOx, CO, CO2, and volatile organic compounds (VOCs) are widely present in industrial flue gas and motor vehicle exhaust. Soot is widely present in diesel vehicle exhaust. The typical contents of these pollutants are shown in Table 1. The annual emissions from 2006 to 2015 of these pollutants in China are shown in the Figure 1. In 2015 year, emissions of SO2, NOx and soot reached 18.59, 18.51, and 15.38 million tons, respectively. Comparing the total emissions, industrial emissions account for nearly 2/3. And the emission of CO2 in 2015 was 9.72 billion tons. Furthermore, the emission of CO and nonmethane volatile organic compounds (NMVOCs) reached 188.00 and 28.43 million tons, respectively. These pollutants are harmful to the atmospheric environment and human health. Emission of SO2 results in acid rain pollution. CO, NO, and VOCs promote the formation of photochemical smog. Soot and VOCs also contribute to haze [1,2]. CO2, N2O, and soot will cause global warming. The global warming potential of N2O is 310 times higher than that of CO2 [3], and soot is theoretically the second largest global warming factor after CO2 due to its radiative forcing effect [1,4]. NOx also causes ozone layer destruction and acid rain [5]. In addition, NOx, CO, and VOCs cause human poisoning. Soot and VOCs cause cancer [6,7]. And soot has a greater impact on the human respiratory system.
There are some ways to remove the above pollutants. Commonly used desulfurization methods include limestone-gypsum technology, rotary spray drying adsorption technology and activated carbon purification technology, which belong to acid-base reaction or adsorption method. However, catalytic reaction is commonly used for the removal of NOx, COx (x = 0, 1, 2), and VOCs, such as catalytic decomposition, reduction, and oxidation. The decomposition of NOx does not require the addition of reducing agents, which is economical and convenient. In particular, the dissociation of N2O is the most commonly used removal method in the production of adipic acid and nitric acid [3]. Selective catalytic reduction (SCR) of NOx by NH3 is widely used in industrial flue gas and motor vehicle exhaust purification. Catalytic oxidation is commonly used to remove multi-pollutants, such as soot, CO, and VOCs, in motor vehicle exhaust and industrial exhaust gas for removal pollutant [11,12]. The preferential oxidation of CO (PROX) is used to purify H2 for fuel cells [13]. Water-gas shift (WGS), reverse water-gas shift (RWGS), partial oxidation of VOCs and the reforming of VOCs are some effective reuse methods. WGS is an important chemical reaction to produce and purify H2, and WGS can adjust the H2/CO ratio for Fischer-Tropsch synthesis [14]. RWGS uses CO2 as a raw material to produce CO, and further produce other chemical raw materials through Fischer-Tropsch synthesis [15]. Partial oxidation of VOCs and reforming of VOCs are the important ways to produce chemical raw materials. For example, partially oxidizing VOCs will produce cyclopropane and hydrogen [16,17]. The CO2 reforming of methane is considered the most popular way to produce syngas [18]. And the steam reforming of methanol is a recognized H2 production process [19].
Removal and reuse methods for these pollutants are often applied in separate modular units existing in industrial flue gas and motor vehicle exhaust. Although the role of potassium species (marked as K) has been found in many reactions, systematic summary and analysis are lacking. The use of the difference in the oxidation-reduction properties of pollutants for coupling control has been favored by more and more researchers [20,21]. Systematic summary and analysis are beneficial to provide K with more favorable support in the catalytic control of multiple pollutants, and the research direction can be better discovered.
In this work, the recent research progress of K on the catalytic reaction process and the properties of catalysts related to the removal and reuse of NOx, COx (x = 0, 1, 2), and VOCs are reviewed. Finally, the connections between the commonly used metals and the action mechanism with pollutants are established. The factors affecting the modification process are also summarized. And the future research direction is prospected.

2. The Removal of NOx

The removal of NOx has been widely studied, including NO decomposition, NO reduction, NO oxidation, and N2O decomposition. And NO reduction and N2O decomposition have been widely applied in industrial flue gas [22].

2.1. N2O Decomposition

The N2O decomposition is the process of adsorbing N2O by catalyst, breaking N–O bond and decomposing N2O into N2 and O2, as shown in Equation (1). This is the most commonly used method for N2O treatment in adipic acid and nitric acid plants. The commonly used catalysts include Co-based and Cu-based catalysts, and the reaction conditions are shown in Table 2. The reaction consists of four steps, the adsorption of N2O on the catalyst, the breaking of N–O bond to form N(ads) and O(ads), the adjacent N(ads) and O(ads) formed on the surface combine with each other to form N2 and O2, respectively, and the desorption of N2 and O2 from the surface.
N 2 O N 2 + 1 2 O 2 .

2.1.1. Promotion Effects of K on Reaction Process

K has a promotion effect on the N2O decomposition [23,24]. K promotes the adsorption and dissociation of N2O. Some researchers have proposed that the electronic actions of alkaline cations enhance the bonds of adsorbates (such as N2O, NO) to accept metal electrons, and then enhance the adsorption of N2O on Pt/Al2O3 and improve the dissociation activity of N2O [25]. K also promotes the formation of oxygen vacancies on TiO2 carrier [26,27]. Moreover, oxygen vacancy on the surface of catalyst is considered to promote N2O adsorption [3,28]. The effect of K promoter on the N2O decomposition is also shown in promoting the dissociation of N-O bond, the limiting step of N2O decomposition [29], through donating electrons [30] and promoting the enrichment of surface electrons [31,32,33].
The addition of K plays a positive role on the dissociation and desorption of O2 in the process of N2O decomposition. Adding K to the Co3O4 catalyst promotes the oxidation regeneration of Co2+ by oxygen dissociated from N2O [34], since K promotes the dissociation of O2. Through temperature-programmed reductions with hydrogen (H2-TPR) test, it is found that the addition of K to Co-Mn-Al catalyst improves the oxygen migration onto the catalyst surface and the desorption ability of O2 [35]. In addition, K promotes the desorption of O2 by improving the reducibility of the catalyst [36,37,38], such as K, promotes the transfer of Ir0 to Ir+ [39]. The promotion effect of K is also shown in accelerating the recovery of active sites in the redox process, and then the K has a promotion effect on the desorption of O2. These two factors together increase the catalytic dissociation activity of N2O.

2.1.2. Promotion Effects of K on Catalyst Properties

In addition to the reaction process, K affects the electronic properties of the catalyst, such as ionization potential and work function, as well as redox properties. The effect of alkali metals on donated electrons to improve the activity of N2O decomposition has been found on a variety of catalysts, such as Co3O4 [40], Co3O4-CeO2 [41], Cu-Co spinels [42], Co-Zn-Ce mixed oxides [43], CuxCo3-xO4 [44], CoAl hydrotalcites [45], NiO [46], and NiAl hydrotalcites [47]. Among them, the excellent promotion effect of K on the decomposition activity of N2O on Co-based catalysts has been confirmed by many researchers [48,49,50,51]. The promotion effect of K is related to their ionization potential and transferring charge to the catalyst, which leads to the decrease of the binding energy [29]. X-ray photoelectron spectroscopy (XPS) results show that the addition of K increases the electron density of the catalyst and promotes the transition of metal electronic state from high valence state to low valence state on Co-based catalysts [52]. It can also be explained that basic cations (K+) provide electrons to surrounding oxygen anions (O2-), and further transfer charge to Co, Mn, and Al [34].
It is found that the addition of K2O reduces the bond energy on Mg/Zn-Ce-Co catalyst, and increases the electron cloud density near Co and promotes the catalytic activity [43]. A large number of studies have shown that the removal activity of N2O is related to the surface work function [53,54,55,56]. Due to its low ionization potential, K significantly reduces the work function of the catalyst by forming a Kn+-On- dipole moment [35], provides electrons to the N2O molecule through the catalyst surface, and activates the adsorption and dissociation of N2O [41].
K has an effect on the redox properties of the catalyst [57]. On the Co-based catalysts, XPS results show that the addition of K promotes the conversion of Co3+ to Co2+, changes the redox characteristics, and thus improves the catalytic activity [43]. The improvement of redox properties of Co2+/Co3+ by K is also found by other researchers [40]. The dissociation of N2O is considered to be the redox mechanism [33,58,59,60,61,62], such as the single-site mechanism shown in Figure 2 [34] and the double-sites mechanism shown in Figure 3 [63]. The improvement of reducibility will significantly promote the dissociation of N2O.
The effect of K on increasing the content of active sites and improving the stability of the catalyst is also studied. The researchers proposed that the good redox cycle of KxOy to KxOy+1 shows an efficient N2O catalytic dissociation activity with the K as an active site on activated carbon [64]. The addition of K to Co3O4 catalyst promotes the stability of Co2+ and the reduction of Co by promoting the desorption of O2 [34]. Moreover, on Ir/Al2O3 catalyst, K improves the resistance to O2 poisoning and improves the stability of N2O decomposition in the presence of O2 [65]. In addition, K is also found to improve the stability in O2 and H2O atmosphere on Zn0.4Co2.6O4/Al2O3 [45]. This may be attributed to the fact that K promotes the desorption of O2 and inhibits the adsorption of O2 on the catalyst surface, thus promoting the stability of the catalyst under aerobic conditions.

2.1.3. Influencing Factors on Promotion Effect

The promotion effect of K on the N2O decomposition is affected by many factors, such as the type and content of alkali metal, preparation method, precursor, and the calcination temperature and time [57,66]. When alkali metal is added to Co4MnAlOx catalyst, the catalytic activity of N2O decomposition increases with the increase of alkali metal ion radius, and the order of promoter activity is Cs > Rb > K > Na > Li [29]. The same results were also found on Co-Ce composite oxide catalysts [41], as shown in Figure 4.
The promotion effect of K is related to the doping amount. When the K contents of 0.3–3.1 wt% are doped on Co-Mn-Al catalyst, the promotion effect increases with the increase of K content, and the promotion effect is the best when the K content is 1.8 wt%. After that, the promotion effect is weakened and the activity is inhibited when the K content is up to 3.1 wt% [67]. The volcanic relationship between the promotion effect and the content is also found on other Co-based catalysts [34].
The preparation methods and precursors greatly affect the activity on K-promoted Co3O4 [56], especially the reduction degree of Co3+ to Co2+ at low temperature. In the comparison of various preparation methods of impregnation, homogeneous precipitation, combustion with glycine, gradual oxidation, and hydrothermal, the catalyst prepared by impregnation method obtains the highest reduction degree from Co3+ to Co2+ and shows the best activity.
The precursors of commonly used K include K2CO3, KNO3, CH3COOK, KOH, etc., as shown in Figure 5 [35,68]. The promotion of activity on Co3O4 has the following relationship: K2CO3 > KOH > KNO3 ≈ KHCO3 > CH3OOK > > K2SO4 ≈ KCl. When K2CO3 is used as a precursor, the lowest light off temperature and the highest reaction rate are obtained with the lowest work function [69]. The poor performance of K2SO4 and KCl is due to the adsorption of SO42- and Cl- on the surface defects, which hinders the decomposition of N2O.
The promotion effect first increases and then decreases with the increase of calcination temperature from 500 °C and 900 °C on Co-Al catalysts, with the best promotion effect appearing at 700 °C [45]. Moreover. the promotion effect of 4 h calcination is higher than that of 12 h. Too long of a calcination time will cause the sintering of the catalyst and affect the promotion effect.
In summary, K affects the adsorption and dissociation properties of the catalyst and then accelerates the catalytic dissociation process of N2O. It affects the ionization potential, work function and redox property of the catalyst, and increases the content of active sites to improve the activity. The stability of the catalyst is improved by promoting the desorption of O2 and stabilizing the active center, similar with the effect on reaction process and the properties of catalyst. The promotion effect of K on N2O decomposition is mainly found on Co-based catalysts, and other catalysts needs to be more studied.
Table 2. Summary of catalysts and reaction conditions for N2O decomposition promoted by alkali metals.
Table 2. Summary of catalysts and reaction conditions for N2O decomposition promoted by alkali metals.
CatalystsPromoterContent (wt%)Reaction ConditionsReferences
Temperature Range (°C)T50 a (°C)N2O (ppm)O2 (%)H2O (%)GHSV b or WHSV c
Co3O4Li, Na, K, Cs0.2200–6003101500317000 h−1[69]
Co3O4K0.01–0.1 d100–500160500022.50.3 g s mL−1[34]
Co3O4-CeO2K2150–4002151000\\0.2 g s mL−1[66]
Co3O4-CeO2Li, Na, K, Rb, Cs0.02–0.07 d150–4002251000\\0.2 g s mL−1[41]
CoxCu3-xO4Na, K,0.005–0.05 d250–6502801000\\0.2 g s mL−1[44]
CoAl2O4Li, Na, K0.04 d200–50029050042.60.12 g s mL−1[45]
Co3O4Li, Na, K, Cs0.015 d450–5505400.976%\\2 g s mL−1[49]
Co3O4Li, Na, K, Rb, Cs0.035 d450–7005701000\\0.5 g s mL−1[52]
Co-Mn-AlLi, Na, K0.3–1.8300–4503300.1%\\0.06 g s mL−1[50]
Co-Mn-AlK0.2–3300–4503500.1%\\40,380 h−1[51]
Co4MnAlOxLi, Na, K, Rb, Cs0.2–3.4300–4503250.1%\\0.06 g s mL−1[29]
Co3O4K0.04–0.1 d200–500260500022.50.3 g s mL−1[56]
Co-Mn-AlNa, K0.5–2.5300–4503300.1%\\0.06 g s mL−1[70]
Co-Mn-AlK0.3–3.1300–4503400.1%\\0.06 g s mL−1[67]
Co3O4Cs0.4–13.750–6001105%\\7000 h−1[68]
Co2.6Zn0.4O4/α-Al2O3K0.15100–6002805%\\7000 h−1[71]
CuO-CeO2Cs0.6–4.8350–60042510002\40,000 h−1[23]
Cu0.8Co0.2Co2O4Na, K, Cs0.05 d300–5003402%4\0.43 g s mL−1[42]
Y2O3-Co3O4K0.32200–4002851%28.212,700 h−1[24]
Zn-Ce-Co3O4Na, K2200–6004503000\\24,000 h−1[43]
NiAl mixed oxidesK0.1300–5003552%48.80.43 g s mL−1[47]
ACNa, K5–20200–6002803000\\0.16 g s mL−1[30]
ACK5–20200–6002853000\\0.16 g s mL−1[64]
a T50 represents the temperature when the efficiency is 50%. b GHSV means gaseous hourly space velocity (h−1). c WHSV means weight hourly space velocity (g s mL−1). d The molar ratio of alkali metal to active center.

2.2. NO Decomposition

For the NO decomposition, similar with the N2O decomposition, NO is adsorbed on the catalyst and decomposed into N2 and O2 due to the catalytic breaking of N-O bond, as shown in Equation (2). NO decomposition is an economical and convenient method to remove NO without using reducing agent. The catalysts commonly used are Co-based catalysts, and the reaction conditions are shown in Table 3. The reaction includes three steps, the adsorption of NO on the catalyst, breaking of N-O bond in NO to form N2 and O2, and the desorption of N2 and O2 on the catalyst.
2 N O N 2 + O 2 .

2.2.1. Promotion Effects of K on Reaction Process

K affects reaction process of NO decomposition, such as the adsorption of NO, the reaction pathway, and the desorption of O2 [72]. The alkaline properties of K affect the NO adsorption and NOx storage, and promote the catalytic activity of NO decomposition [73]. This is consistent with the mechanism of K in N2O decomposition reaction. Basic cations enhance the bond strength of adsorbates which can accept metal electrons by electronic interaction [25]. Some researchers also believe that the K regulates the basicity of catalyst surface, thus promoting the adsorption of NO [74,75].
K has an effect on changing the reaction pathway of NO decomposition. On the Co3O4 catalyst, it is found that NO is first converted to NO2- adsorbed on K, then transfers to the interface between K and Co3O4, and reacts with adsorbed NO to form N2 under the catalysis of Co [76]. The K-Co interface formed after doping with K has a higher catalytic activity of NO decomposition [77].
The addition of K also promotes the desorption of O2. Haneda [77] proposed that the desorption of O2 is the limiting step of NO decomposition. The redox property of O2 dissociation is improved by K on Co-Mn-Al catalyst [73]. Moreover, K increases the electron density of Co and weakens the strength of Co-O bond, which is beneficial to promote the desorption of O2 on Co2+ [45,52]. In addition, K improves the dissociation activity of NO under O2, CO2 and H2O atmospheres over Co-Mn-Al catalyst due to the improvement of basicity and reducibility [78].

2.2.2. Promotion Effects of K on Catalyst Properties

K affects the properties of NO decomposition catalysts, including redox properties, basicity, electronic properties and stability. On the K-promoted Co-Mn-Al catalyst, it is found that K increases the reducibility and basicity of the catalyst [73]. Consistent with the role of K in the N2O decomposition, due to its low ionization potential, K can transfer charge to the transition metal cation, resulting in a dipole that can produce an electric field gradient on the catalyst surface [67]. K increases the specific activity per unit area of Co3O4 catalyst [52], and the effect of K on increasing the content of active Co2+ sites has also been found by other researchers [49], in spite of weakening the strength of Co-O bond and promoting the formation of low-valent cobalt species [52].
As K loaded on Co-Mn-Al catalyst, the changes well reflect the effect of K on the NO decomposition [73], as shown in Figure 6. K modification makes the surface of catalyst rich in potassium species, which increases the number of basic sites, especially the medium and strong basic sites. K also promotes the formation of active Co2+ sites and the adsorption of NO. The increase of catalyst basicity and reducibility further promotes the catalytic dissociation activity of NO and the desorption activity of O2.

2.2.3. Influencing Factors on Promotion Effect

The promotion effect of K on NO decomposition reaction is affected by many factors, such as type and content of alkali metals, calcination, and preparation method. It is found that the order of promotion effect is, K > Na > Rb > Cs, by adding Na, K, Rb, and Cs to Co3O4 catalysts [76], and K shows an excellent promotion effect on other Co-based catalysts [52]. This is different from the findings in the N2O decomposition. The promotion effect is affected by the K content. When the Co-based catalyst with Co/Al mole ratio of 3 is supported with K of 0.04–0.12 K/Co molar ratio, the K content of 0.08 shows the best promotion effect [45]. Some researchers have proposed that the catalytic dissociation process of NO on the catalyst surface is related to the surface coverage, crystal phase, and surface defect content [79]. K improves the surface NO adsorption, and increases its coverage and dissociation activity, but too high NO coverage will inhibit the dissociation of NO due to the lack of vacancies to desorb O2.
K-modified Co-based catalysts are greatly affected by calcination temperature and calcination time [45,71]. The calcination temperature mainly affects the migration and redistribution of K. The higher the calcination temperature, the more stable K is [73]. However, K volatilizes from transition metal oxides above 500 °C, which may affect the long-term stability of the catalyst [80]. Calcination at high temperature also affects the surface state of K [81]. K migrates from carrier, acting as K reservoir, to spinel nanocrystals and then forms KxCoO2 coating at high calcination temperature [82], leading to the disappearance of the strong beneficial effect of K.
The preparation method of K-promoted catalyst has an effect on the activity and stability of NO decomposition. The co-precipitation of metal nitrates by a Na2CO3/NaOH and subsequent impregnation KNO3 leads to a higher activity and stability of the catalyst than the other preparation methods, such as the calcination of corresponding metal nitrates or the impregnation of the calcined Co-Mn-Al precursor with a solution of KNO3 [80]. In the preparation of K-promoted Co-Mn-Al catalyst, the catalyst prepared by coprecipitation of transition metal nitrate and K2CO3/KOH shows a good catalytic activity and stability for the NO decomposition. Addition of K has a positive effect on the stability of the catalyst because K acts as a storage for the catalyst when the K is desorbed from the catalyst surface [72]. Some researchers have also proposed the model shown in Figure 7 for the improvement of the stability of the K-doped catalyst. The more stable KFeO2 is used as the outer layer for the reaction, and K2Fe22O34 is the K reservoir, which constantly supplies the surface K content [81].
K shows promotion effects on both N2O decomposition and NO decomposition. On the two types of catalysts, K has the same effect on redox property, increasing the electron density of the active center and weakening the bond energy. Taking the Co-based catalysts commonly used in both reactions as examples, the addition of K promotes the conversion of Co3+ to Co2+ and the recovery of the active center during the reaction. There are also some different effects of K in the two kinds of reactions. In the N2O decomposition, the enhanced adsorption of N2O by K comes from the electron-donating action, while in the NO decomposition, the enhanced adsorption of NO by K comes not only from the electron-donating action, but also from the alkaline action. In addition, K itself can be used as the active site for N2O catalytic dissociation on K/AC catalyst in N2O decomposition, which has not been studied and discovered in NO decomposition. K changes the reaction pathway by forming the NO2- on K in NO decomposition, which is not discovered in N2O decomposition. The promotion effect is consistent with the ion radius of the alkali metal in N2O decomposition, while K shows the best promotion effect in NO decomposition. Furthermore, research on the influence of precursor is lacking.
Table 3. Summary of catalysts and reaction conditions for NO decomposition promoted by alkali metals.
Table 3. Summary of catalysts and reaction conditions for NO decomposition promoted by alkali metals.
CatalystsPromoterContent (wt%)Reaction ConditionsReferences
Temperature Range (°C)T50 a (°C)N2O (ppm)GHSV b or WHSV c
Co-Mn-AlK0.6–18.9650–70065010000.6 g s mL−1[72]
Co-Mn-AlK1–4560–65065010000.6 g s mL−1[73]
Co3O4Li, Na, K, Rb, Cs0–0.1 d450–70058010000.5 g s mL−1[52]
Co3O4Na, K, Rb, Cs0.035 d450–60060010000.1 g s mL−1[76]
Co3O4K0.9–3400–650<65097002100 h−1[83]
Co3O4Na0–0.091450–7006200.9762 g s mL−1[49]
Co-Mg-Mn-AlK255–70066010000.6 g s mL−1[84]
Co-Mn-AlK, Cs1.5–4612–650\10000.6 g s mL−1[80]
Ce-Mn Li, Na, K, Cs0–10800>80060001 g s mL−1[85]
a T50 represents the temperature when the efficiency is 50%. b GHSV means gaseous hourly space velocity (h−1). c WHSV means weight hourly space velocity (g s mL−1). d The molar ratio of alkali metal to active center.

2.3. NO Reduction

NO reduction is generally a process of selective catalytic reduction of NO by NH3 to form N2 and H2O through the catalyst, which is shortly called NH3-SCR, as shown in Equations (3) and (4). SCR is the most common technology for NOx removal from industrial flue gas and motor vehicle exhaust. The commonly used catalysts are V-based, Mn-based, and Cu-based catalysts and the reaction conditions are shown in Table 4.
4 N O + 4 N H 3 + O 2 4 N 2 + 6 H 2 O ,
N O + N O 2 + 2 N H 3 2 N 2 + 3 H 2 O .

2.3.1. Promotion Effects of K on Reaction Process and Catalyst Properties

The promotion effect of K on the NH3-SCR has been reported, such as the adsorption and reaction pathway of NO and NH3. The promotion effect on the basicity, electron donor properties and active center stability of NH3-SCR catalyst is also found. At present, the widely accepted NH3-SCR mechanisms are the Langmuir–Hinshelwood (L-H) mechanism and Eley–Rideal (E-R) mechanism. The L-H mechanism is that NH3 and NO are adsorbed on the catalyst surface, and the adsorbed NH3 reacts with adsorbed NO to produce N2 and H2O. The E-R mechanism is that NH3 is adsorbed on the catalyst surface and the adsorbed NH3 reacts with gaseous NO to produce N2 and H2O. Different from the traditional mechanism, gaseous NH3 reacts with adsorbed NO2 to complete SCR reaction, which is called fast SCR reaction [86]. Some researchers have proposed that K provides alkaline site to adsorb NO2 in fast SCR. And the doping of K promotes the chemisorption of NO through alkaline and electron donor interaction, and forms a new reaction pathway on the catalyst, thus accelerating the SCR reaction [87]. Moreover, the catalytic dissociation activity of K on N2O promotes the decomposition of N2O during SCR reaction to form N2, which improves the selectivity of N2. K increases the activity of unsaturated Mn cations in α-MnO2 catalysts, promoting the chemisorption and activation of NH3 [88]. Alkali metal cations stabilize the metastable active sites of Fe, Au, etc., thus promoting the catalytic activity [89,90]. K also affects the stability of catalyst. The addition of K on Cu/SAPO-34 decreases the likelihood of water attack due to the decrease in the density of Si-O(H)-Al bonds, which significantly improves the stability in 3 vol% H2O atmosphere [91].
Besides NH3-SCR, selective catalytic reduction by CxHy (HC-SCR) or H2 (H2-SCR) and the reduction of NO by CO (NO-CO reaction) have also been promoted by K. For the reduction of NO by propene (C3H6) on the Pt/γ-Al2O3 catalyst, K promotes the adsorption and dissociation of electronegative NO and inhibits the adsorption of electropositive C3H6, thus adjusting the adsorption of reactants [92]. Moreover, K strengthens the Pt-NO bond and weakens the N-O bond, thereby increasing the HC-SCR activity [93]. K also shows the same effect in the NO-CO reaction on Pt/γ-Al2O3 catalyst [94]. K improves the dispersion of active centers on Pd/K2O-6TiO [95] and Ag/γ-Al2O3 [96] in H2-SCR. In addition, the content of the intermediate species, Pd0-NO, are also increased by K on Pd/K2O-6TiO2 [95]. On the CuO/AC catalyst, K promotes the re-oxidation of Cu+ to Cu2+ and the formation of surface active carbon sites by weakening the C-C bond, which promotes the redox cycle of active sites [97]. K has a promotion effect on the carrier. The modification of K to TiO2 with loading Au have been found forming K2Ti8O17, which promotes the formation of isocyanide (-NCO) [98]. -NCO is an important reaction intermediate, which is considered to be a key to the high activity and selectivity of NO-CO reaction [99].
K also has a promotion effect on selectivity. K promotes the dissociation of NO, increases the content of Nads on the Pt/γ-Al2O3, and relatively reduces the content of NOads [92], thereby promoting the formation of N2, as shown in Equation (5), and weakening the formation of N2O, as shown in Equation (6). This effect is also found on Pd-based catalysts [100].
N a d s + N a d s N 2 g ,
N a d s + N O a d s N 2 O g .
K promotes the stability of the catalyst. Sulfur resistance of catalyst is enhanced by adding alkali metals on Ag/γ-Al2O3 catalyst due to the increase of intermediates [96]. The SO2-TPD results show that the addition of K inhibits the accumulation of sulfur compounds on Ag/Al2O3 in HC-SCR, thereby increasing the stability in SO2 atmosphere [101]. In the reduction of NO by CO and naphthalene reaction on Cu/Ce/TiO2-SiO2, K reduces the oxidation of naphthalene, thereby promoting the reduction of NO [102].

2.3.2. Influencing Factors on Promotion Effect

The promotion effect is affected by the type of alkali metal, the content and the reaction temperature. The effects of Li, K, Rb, and Cs in HC-SCR over Pt/γ-Al2O3 catalyst are compared [93], and Rb shows the best promotion effect. In the reduction of NO by propene over Pd/Y2O3-ZrO2 catalyst, Na shows the best promotion effect over Li, Na, K, and Cs [100], while Rb shows a better promotion effect than K and Cs in the NO-CO reaction over Pt/γ-Al2O3 catalyst, due to the suitable ion size of Rb [94]. There are different optimal types of alkali metals for different active sites in NO reduction.
There is a volcanic relationship between the content of K and the promotion effect. The addition of 4.2 wt% K to Pt/γ-Al2O3 catalyst performs the best promotion effect among the content range of 2.6–10.4 wt% in HC-SCR [92]. The alkali metal contents of 0.02–0.1 wt% over Pd/Y2O3-ZrO2 catalyst are studied in HC-SCR and NO-CO reaction [100] with the best promotion effect at 0.02 wt% content. The alkali metal contents of 0.5–2 wt% are studied on Ag/γ-Al2O3 catalyst in HC-SCR [96], with the best promotion effect at 1 wt% alkali metal content.
In the H2-SCR reaction over Pd/K2O-6TiO2, the increase of reaction temperature reduces the reaction activity and selectivity due to the promotion in the decomposition of nitrate adsorbed by K to form NO2 and N2O [95].
K plays the same role in the both reaction of SCR and NO decomposition for the NOx removal. Although the catalyst active centers promoted by K in the two reactions are inconsistent, K shows a similar promotion effect. The electron-donating properties of K promote the adsorption of NO, the stability of metastable active sites, activity and stability. And K has the effect of changing the reaction pathway in both reactions. The effect of the content on the K promotion effect is found to be volcanic in both reactions, indicating that the effects of K are both positive and negative. A proper amount of K promotes the adsorption of NO, while too much K will destroy the adsorption balance and cover the active sites. In the NO reaction, K is found to promote the dispersion of Ag and Pd, which is an undiscovered role on Co-based catalysts in the NO decomposition. There are few researches on factors affecting the promotion of K in NO reduction, especially the precursor. And the research on stability in SO2 and H2O on K-promoted catalysts is lacking.
Table 4. Summary of catalysts and reaction conditions for NO reduction promoted by alkali metals.
Table 4. Summary of catalysts and reaction conditions for NO reduction promoted by alkali metals.
CatalystsPromoterContent (wt%)Reaction ConditionsReferences
Temperature Range (°C)T50 a (°C)NO (ppm)ReductantOthersGHSV b or WHSV c
Pt/γ-Al2O3Li, K, Rb, Cs1.9–15.5177–52730010001000 ppm C3H6\0.006 g s mL−1[103]
Pt/γ-Al2O3Na2.6–10.4200–50030010001000 ppm C3H6\0.006 g s mL−1[92]
Pt/γ-Al2O3K, Rb, Cs1.9–15.5100–50032010001000 ppm CO\0.006 g s mL−1[94]
Pd/K2O-6TiO2K\75–2759510005000 ppm H25% O260000 h−1[95]
Pd/Y2O3-ZrO2Na0.017–0.102300–45036280008000 ppm C3H6\0.00375 g smL−1[104]
Pd/γ-Al2O3Na3.5–7180–35027510001067 ppm C3H6
7000 ppm CO
7800 ppm O20.015 g s mL−1[105]
Ag/Al2O3Li, Na, K, Cs0.5–1200–50035010001000 ppm C3H65% O20.09 g s mL−1[96]
Ag/Al2O3Na, K, Cs2450–7005455004000 ppm CH480 ppm SO29000 h−1[101]
CuO/ACK0–10150–4503172000AC\20,000 h−1[97]
a T50 represents the temperature when the efficiency is 50%. b GHSV means gaseous hourly space velocity (h−1). c WHSV means weight hourly space velocity (g s mL−1).

2.4. NO Oxidation

For the NO oxidation, NO and O2 are first adsorbed on the catalyst surface and then react to form NO2, as shown in Equation (7). The commonly used catalysts are Pt-based, Mn-based, and Co-based catalysts and the reaction conditions are shown in Table 5.
2 N O + O 2 2 N O 2 .

2.4.1. Promotion Effects of K on Reaction Process

K promotes reaction process of NO oxidation, such as the adsorption of NO and activation of O2. K enhances the strength of Pt-NO bond and promotes the formation of adsorbed N species (nitrites/nitrates) [106], which promote the oxidation of NO into NO2 on Pt/Al2O3 [6]. In addition, K provides the adsorption site for NO thereby reduces the competitive adsorption since O2 is a stronger electron acceptor than NO which is preferentially adsorbed on the Pt site [107]. K has also been found to promote the dissociation of O2 on the 2D hexagonal boron nitride [108], which is considered to be the limiting step in the oxidation reaction [109].

2.4.2. Promotion Effects of K on Catalyst Properties

K changes the properties of catalysts, such as acid-base properties and redox properties. K reduces the crystallinity of the catalyst, improves the dispersion of surface particles, and changes adsorption characteristics on Mn-CoOx [110]. In addition, K improves the dispersion and the ratio of Mn4+/Mn3+ and Co2+/Co3+, and promotes the chemisorption of O2 and NO. The increasing of the ratio of Co2+/Co3+ by adding K is also found on Co/KxTi2O5 [103]. Moreover, Fourier transform infrared (FTIR) results show that K in the carrier TiO2 promotes the formation of chemisorbed species of NOx on the catalyst surface.

2.4.3. Influencing Factors on Promotion Effect

The promotion of K on NO oxidation is affected by the precursor, content and existing states of K. Among the three kinds of precursors of KOH, K2CO3 and KNO3 on Mn-CoOx, KOH shows the best activity, which is consistent with the alkalinity. It is found that the catalyst with 10 wt% K in the range of 5–20 wt% K contents shows the best catalytic activity for NO oxidation over Mn-CoOx catalyst [110], while the Co3O4 catalyst with 0.1 wt% K in the range of the 0.05–0.2 wt% K contents shows the best catalytic activity for NO oxidation [111]. On the Co/KxTi2O5 catalysts, 2.15–15.28 wt% K contents are studied [103], and it is found that the NO oxidation activity increases as the increasing of K content. There are different suitable K contents on different catalysts. For the existing states of K, two forms of K are found on Co3O4 catalysts, free K and stable K [111]. Free K species, including carbonate and nitrate, have high mobility and easily cover active sites, which is adverse for the oxidation of NO. Stable K species closely relate to the activity serves as electronic and structural modifiers, which is conducive to the formation of O/O2- species, oxygen vacancies, more Co3+ sites, and then is conducive to the oxidation of NO.
NO decomposition and NO oxidation are two opposite reaction pathways of NO to form N2 or NO2. K promotes the adsorption of NO in the both reactions, but the mechanism is different. In the NO decomposition, K promotes the formation and stability of the low valence metal, while in the NO oxidation, K regulates the ratio of metal valence to favor the reaction, and simultaneously increases the dispersion of the active centers, thus improving the redox performance and the oxidation activity. Especially on Co-based catalysts, the promotion of K on the formation of Co2+ or Co3+ requires a clearer analysis. Moreover, K plays more promoting roles in NO decomposition than that in the NO oxidation, such as the activation of NO and the increase of the specific surface area of the catalyst. In NO oxidation, the existing states of K are proposed, which is worthy of attention in the research. Moreover, the optimal K precursor is KOH in NO oxidation and K2CO3 in N2O decomposition, showing that different reactions require different catalyst properties.
Table 5. Summary of catalysts and reaction conditions for NO oxidation promoted by alkali metals.
Table 5. Summary of catalysts and reaction conditions for NO oxidation promoted by alkali metals.
CatalystsPromoterContent (wt%)Reaction ConditionsReferences
Temperature Range (°C)T50a (°C)NO (ppm)O2 (%)GHSV b or
WHSV c
Co/KxTi2O5K2.15–15.28200–42027570010120,000 h−1[103]
Co3O4K0.05–0.2200–45024050080.27 g s mL−1[111]
Mn-CoOxK550–25085500530,000 h−1[110]
Ru/K-OMS-2K\27–5272701000835,000 h−1[112]
a T50 represents the temperature when the efficiency is 50%. b GHSV means gaseous hourly space velocity (h−1). c WHSV means weight hourly space velocity (g s mL−1).

3. The Removal and Reuse of COx

K has been widely studied for promoting removal and reuse of COx (x = 0, 1, 2), such as catalytic oxidation to remove soot and CO, and water-gas shift (WGS) reaction and reverse water-gas shift (RWGS) reaction to reuse CO and CO2.

3.1. Soot Oxidation

Soot oxidation is oxidizing soot particles by oxidizing species on catalysts to form CO2, as shown in Equation (8). The oxidation of soot is the most common method for soot removal in motor vehicle exhaust. The commonly used catalysts are Ce-based, Cu-based, Mn-based and Co-based catalysts. The reaction conditions are shown in Table 6. The reaction includes three steps, the adsorption of soot and oxidizing species on the catalyst, the reaction between soot and oxidizing species to produce CO2, and the desorption of CO2. The soot also can be oxidized by NO due to its existence in motor vehicle exhaust.
C + O 2 C O 2 .

3.1.1. Promotion Effects of K on Reaction Process

K shows a promotion effect in soot oxidation reaction [113], including the adsorption of O2, the contact between soot and catalyst, the reaction pathway and the selectivity of CO2. The addition of K to Ce-based catalysts promotes the adsorption of O2 and the formation of surface carbon complexes which can react with soot, thus improving the activity of soot oxidation [114]. Adding K into Mg-Al catalyst shows that surface oxygen on the K active site is reactive oxygen species in the reaction [115]. In addition, K also promotes the dissociation of O2. The XPS and in-situ diffuse reflectance infrared Fourier transform (in-situ DRIFTS) results show that K on the catalyst surface reacts with O2 to form peroxides and superoxide-type substances with high oxygen content which can participate in the soot oxidation process. This is also considered to be the main reason for the increased activity of K promoted catalysts [116]. The low melting point of KNO3 greatly improves the contact between catalyst and soot on Cu/CeO2 [117]. The existence of K also increases the number of structural defects on soot and improves the contact between the catalyst and soot particles, thus increasing the utilization of active sites [118]. The free K species are found on Co3O4, in the forms of carbonates and nitrates with high mobility, which improve the contact state of catalyst-soot and significantly accelerate soot oxidation [111].
The effect of K on the reaction pathway has also been proposed. The process of soot catalytic oxidation could be changed by the molten phase KNO3. The molten phase KNO3 not only promotes solid-liquid contact but also promotes the fragmentation of soot particles and creates a new reaction pathway for soot oxidation, as shown in Figure 8. Soot is oxidized by the oxygen in the molten phase nitrate, and the gaseous O2 is used to supply the oxygen consumed in the molten phase nitrate [119]. The oxidation pathway of soot changes from the direct oxidation by gaseous O2 to the indirect oxidation, so soot oxidation can be transferred from high temperature to low temperature. On the K/MgAlO catalyst, active oxygen is formed on K to participate in the oxidation process and the oxidation of soot follows the oxygen overflow mechanism [120]. The active oxygen existing at the K site overflows to the surface of soot particles, reacts with the free carbon to form an intermediate ketene with a C=C=O structure, and then the active oxygen species at the K site are continuously supplemented by gaseous O2 and lattice oxygen, and ketene is further oxidized to CO2, as shown in Figure 9.
In the oxidation of soot by NO and O2 on Mg-Al catalyst, it is found that K provides new active oxygen species and changes the reaction pathway for the oxidation of soot by NO [115]. As shown in Figure 10, NO is first oxidized to nitrites by the active oxygen at the K site. Then the nitrites react with the C on the soot to form the ketene group. Finally, the ketene group is further oxidized to CO2. The same effect is also found on K-modified Co3O4 nanowires monolithic catalysts (KCo-NW), as shown in Figure 11 [121]. K promotes the chemical adsorption of NO and forms chemisorbed NOx species on the surface. Among 200–250 °C, the chemisorbed NOx species on the catalyst surface participate in the soot oxidation reaction. Among 250–330 °C, the surface chemically adsorbed NOx species begin to decompose, and the produced NO2 participates in the soot oxidation reaction. Above 330 °C, soot is oxidized by gaseous NO2 and O2. The addition of K forms new reactive oxygen species, which causes soot oxidation to occur at lower temperature. Moreover, on the CeO2 catalysts, K increases the possibility of multiple reaction pathways by acting as a reaction site, thus improving the oxidation activity of soot, as shown in Figure 12 [118]. After adding K salts, soot oxidation mechanism is no longer merely oxygen transfer mechanism, but also electron transfer mechanism [122]. The effect of K on CO2 adsorption, promoting the formation of CO2 and improving selectivity, is also found on CaO-MgO [123,124].

3.1.2. Promotion Effects of K on Catalyst Properties

Researchers have found that alkali metal additives can be directly used as active centers. K+ is considered to be the active site of soot oxidation on KNbO3 catalyst [125]. K on the KNO3 catalyst in the molten phase is also directly used as the active centers for soot catalytic oxidation [119]. Moreover, the potassium oxides and peroxides are considered to be active species for the reaction of C with O2 [126].
K affects the redox properties of the catalyst. The addition of K enhances the reducibility on CaO-MgO catalyst [116], promotes the formation of large oxygen-containing surface substances on MgO [127], forms new reactive oxygen species, K-Co-O, on Co-MgAlO catalyst [128] and increases the number of oxygen vacancy on Mg-Al catalyst [120], which significantly improves the oxidation activity of soot.
K also promotes the stability of the catalyst. It is found that the sulfur resistance of the MnOx-CeO2 and Cu/CeO2 catalyst is improved by impregnating KNO3 [117,129], which is attributed to the protection of the active center by the basicity of K. The addition of K to the CaO-MgO catalyst reduces the content and stability of carbonates formed on the surface, while the carbonate groups invalidate the active sites and prevent the adsorption and dissociation of O2 [116]. K reduces decomposition temperatures of carbonate on catalyst surfaces is also found [130].

3.1.3. Influencing Factors on Promotion Effect

The promotion effect of K on soot oxidation catalyst is affected by the type of alkali metal, the content and the calcination temperature. In a series of MNbO3 catalysts (M = alkali metal), the promotion effect has the following relationship: KNbO3 > NaNbO3 > RbNbO3 > LiNbO3 [125]. Some researchers found that promotion effect of alkali metals increases with the increase of atomic number, that is, K > Na > Li [131,132,133,134,135,136]. The linear relationship between promotion effect and the content is found on Cu/CeO2 in molar ratio of K/Cu from 0.4 to 2.5 [117] and on Co-MgAlO catalysts among 1.5–10% K [128]. However, the volcanic relationship between promotion effect and the content is found on Ce-based catalysts among 3–13.5% K [114].
The XRD results show that the appropriate increase of calcination temperature increases the crystallinity of KNbO3 catalyst, but too high calcination temperature reduces the specific surface area, and the calcination temperature at 650 °C shows the best activity [125]. The promotion effect is affected by calcination temperature due to the type of salts. K forms KNO3 species at low calcination temperatures (450 °C) and forms K2O species at high calcination temperatures (650 °C) on the Fe/Al2O3 catalysts [6]. Different existing states of K have different effect on activity. KNO3 plays a promoting role in the reaction, while K2O will combine with the product CO2 to form K2CO3 and further decompose to produce basic O2-, which covers the active sites, thereby reducing the soot oxidation activity.
K can act as an active center in soot oxidation, which changes the reaction pathway. This is not found in the NOx removal. The researchers discovered the deep meaning of the influence of calcination temperature, which changed the existing states of K on the surface. However, there are some contradictions in soot oxidation and NOx removal. Although K can promote the adsorption of reactants in both reactions, it is a new discovery in the soot oxidation reaction to promote soot adsorption by constructing defect sites. It is found that K promotes the desorption of O2 on the Co-based catalyst during NOx decomposition, while K promotes the adsorption of O2 on CeO2 during soot oxidation. K promotes the formation of active oxygen species on the catalyst surface in soot oxidation, which is also different from the findings in NOx removal. The difference of K in soot oxidation and NOx removal may be due to the nature of the active center and the atmosphere conditions, and more in-depth research is needed.
Table 6. Summary of catalysts and reaction conditions for soot oxidation catalysts promoted by alkali metals.
Table 6. Summary of catalysts and reaction conditions for soot oxidation catalysts promoted by alkali metals.
CatalystsPromoterContent (wt%)Reaction ConditionsReferences
Temperature Range (°C)Tm a (°C)Soot (g Soot/g Catalyst)O2 (%)H2O (%)NO (ppm)OthersGHSV b or WHSV c
MnOx-CeO2K0.08–0.67 d300–7004400.051031000\0.043 g s mL−1[129]
CeO2K4–14300–6503500.056\\\0.024 g s mL−1[118]
CeO2K3–13.5300–5003600.056\\\0.00375 g s mL−1[114]
CeO2-ZrO2K8400–6504300.110\\1000ppm NO0.0132 g s mL−1[137]
Cu/CeO2K2–5300–5003150.110\1000\0.0132 g s mL−1[117]
Cu/Al
Co/Al
V/Al
K0.5–1 d250–6505500.25\500\≈1270 h−1[113]
Co-MgAlOK1.5–1050–3003480.05air\\[128]
CaO-MgOLi, Na, K5.4250–7504300.25air0.0025 g s mL−1[116]
MgOK2.3350–7004210.25air0.0025 g s mL−1[127]
MgAlOK0.1450–500217e\10\0.1%0.5% CO, 0.05% C3H620,000 h−1[120]
MgAlOK≈1.9–3.9376–7504100.11air\\[138]
Alkaline salts Li, Na, K\250–5003500.510\\\0.138 g s mL−1[119]
alkaline niobatesLi, Na, K, Rb\400–7004700.2512\\\0.0025 g s mL−1[125]
a Tm represents the temperature when the efficiency is maximum. b GHSV means gaseous hourly space velocity (h−1). c WHSV means weight hourly space velocity (g s mL−1). d The molar ratio of alkali metal to active center. e This temperature represents ignition temperature.

3.2. CO Oxidation

CO oxidation is that adsorbed CO is oxidized by O2 to generate CO2, as shown in Equation (9). This is one of most effective ways to reduce CO emissions in industrial flue gas and motor vehicle exhaust. The commonly used catalysts are Pt-based, Cu-based, Au-based, and Rh-based catalysts. The reaction conditions are shown in Table 7. The reaction goes through three steps, the adsorption of CO and O2 on the catalyst, the reaction of adsorbed CO with active oxygen species to form CO2, and the desorption of CO2. This is similar to the catalytic oxidation of soot. In addition, high preferential oxidation activity of CO in H2 is beneficial to the acquisition of high purity hydrogen fuel, which is very important for fuel cells.
2 C O + O 2 2 C O 2 .

3.2.1. Promotion Effects of K on Reaction Process

K favors the adsorption O2, changes the reaction pathway and improves the selectivity of CO2 [1,13,122,139]. K affects the adsorption of CO and O2. On the one hand, K weakens the adsorption of CO and promotes the adsorption of O2. On Pt-based catalysts, the addition of K weakens the interaction between CO and surface Pt atoms, reduces the coverage of CO, and thus improves the chemical adsorption of competitive O2 on Pt atoms [140]. This weakening effect is related to the electronic action of K [141]. The FTIR results on Pt/Al2O3 catalyst also show that the K weakens the relationship between CO and Pt, and changes the adsorption site of CO [142]. And the interaction between CO and Pt becomes weaker with the increase of K content [143]. According to the kinetic study, K decreases the adsorption equilibrium constant of CO and reduces the chemisorption of CO, which promotes the chemisorption of O2 and significantly reduces the reaction barrier between chemisorbed CO and O2 [140]. The doping of K on Ru-based catalysts also inhibits the adsorption of CO, thus promotes the adsorption of O2 through competitive adsorption at lower temperatures [144]. Some researchers have also proposed that the K promotes the adsorption of O2 by changing the binding energy [145,146]. In addition, with the addition of K, the formed K2O provides a site for the adsorption of O2 to promote the adsorption of O2 [147]. The role of K in promoting the dissociation and adsorption of O2 is also found on Ce0.8Zr0.2O2 catalysts [114].
On the other hand, the addition of K to Pt-based catalysts has promoting effect on strong adsorption of CO. Although K reduces the overall CO adsorption content on Pt-based catalysts, the formed K-O-Pt species have strong adsorption to CO and reduces the reaction barrier, thus obtaining higher activity [140]. Moreover, K also improves the adsorption properties of CO and O2 [145,146,148]. The addition of K on Rh/ultra-stable Y zeolite (USY) promotes the adsorption of CO by increasing the formation of linear CO and bridged CO [149], which is consistent with the results observed on Rh/Al2O3 [150]. The addition of K on Au/TiO2 catalyst also promotes the adsorption of CO, and K is considered as serve as anchor sites for adsorbates [151].
The opposite effects of K on the adsorption of CO are worth noting. K promotes the adsorption of CO on the Rh-based and Au-based catalysts by increasing the CO adsorption species, acting as the CO adsorption anchor, forming new species and generating electrostatic interaction between Kδ+ and Oδ-. The difference is that, on the Pt-based and Ru-based catalysts, K weakens the strength of the Pt-C bond by increasing the electron density near the Pt site, thus inhibiting the adsorption of CO. The role of K on Pt-based catalysts is special. The addition of K reduces the adsorption equilibrium constant of CO and improves that of O2. But the new species, K-O-Pt, have higher CO adsorption activity, reduce the energy barrier of CO oxidation reaction, and promote the reaction rate.
K changes the reaction pathway of CO oxidation in the catalytic process. The formed K2O provides a site for the adsorption of O2 on the Pt/Al2O3 [147]. The O2 adsorbed on K2O reacts with the CO adsorbed on the adjacent Pt site to complete the oxidation of CO, and then the reaction pathway between CO and O2 is changed from the competitive adsorption reaction on single active site to the non-competitive adsorption reaction on double active sites. The addition of K on Pt-based [140], Ir-based [152], and Au-based catalysts [151] has been found to change the reaction pathway which significantly reduces the reaction barrier and the disproportionation reaction of CO occurs on Au/TiO2, as shown in Equation (10).
2 C O C a d s + C O 2 , a d s .
K also improves the selectivity of CO2 in the reaction process. On both Pt/Al2O3 and Rh-based catalysts, the addition of K improves the selectivity of CO2 [153,154]. On Ni-based catalysts, as shown in Figure 13, Alkali metals inhibit the formation of carbonyl nickel on the surface, prevent the occur of methanation, and improve the selectivity [155].

3.2.2. Promotion Effects of K on Catalyst Properties

K affects the properties of CO oxidation catalysts, such as promoting the stability of active centers, increasing the number and activity of active centers, improving redox properties, and affecting surface active oxygen. K stabilizes the structure of Au on Au-based catalysts [156] and the metal state of Rh particles by the interaction between K and acidic sites on USY zeolite [157], thereby increasing the CO oxidation activity. K weakens the interaction between Ru and carrier silica, reduces the reduction stability of Ru, and keeps Ru in the state of metal Ru on Ru/SiO2 [144] and stabilizes Cu+ on Cu-based catalysts [158] through the characteristics of electron donor.
K improves the dispersion, quantity and activity of active centers. K is considered to be served as anchor sites for Au on Au/TiO2 catalyst [159]. K can improve the dispersion of Rh on Rh/SiO2 [154], increase the percentage of surface active center Ru0 on Ru/SiO2 [144] and change the surface Rh species to form more active sites on Rh/USY [157]. K also affects the redox property of the active center. The H2-TPR result shows that the addition of K to CuO-CeO2 catalyst decreases the reduction temperature of metal oxides [155], and XPS characterization reveals that the content of Ce3+ significantly increases [18].
K affects the surface reactive oxygen species and promotes the oxidation performance of the catalyst. The addition of K to the Pt/TiO2 [26] and CuO/CeO2 [160] promotes the formation of oxygen vacancy, which enhances the oxygen storage capacity, and increases the number of surface active oxygen species. K increases the proportion of reactive oxygen species, the peroxide and superoxide formed, in the process of O2 adsorption on the surface of Ce0.8Zr0.2O2 [161]. K transfers electrons to the Ce-O bond on Ce-based catalysts, thus increases the reactivity of oxygen, and prevents the reduction of oxygen capacity [137]. K enhances the synergism between Cu and CeO2 on CuO-CeO2 [160]. K also participates in the redox cycle of Ce [114]. The alkalinity of K has a positive effect on catalyst. K can interact with the catalyst on the Pt/Al2O3 and improve the coverage of OH groups on the surface, which participates in the catalytic oxidation of CO [162].
The doping of K promotes the stability of the catalyst. K improves the resistance of CO2 and H2O on Ru/SiO2 [157]. In addition, K prevents carbon deposition on Ni-based catalysts, thus promoting the stability of the catalyst [18].
The preferential catalytic oxidation of CO in H2 atmosphere has also been widely studied. K shows a unique promotion effect on PROX, in other words, K makes H2 in the atmosphere no longer inhibit or even promote the oxidation activity of CO, since K suppresses the competition of H2 on the catalysts. On Rh/USY catalyst, the CO oxidation activity of unmodified catalyst decreases significantly in H2-rich atmosphere, while the activity of K-modified catalyst is not affected by the partial pressure of H2 [157] because K inhibits the oxidation of H2 [163]. The in-situ DRIFT result shows that K enhances the adsorption of bridged CO, so it is difficult to activate H2 on Rh/USY, which leads to the high PROX activity. Similarly, with the addition of K on Rh-based catalysts, the electron donor effect of K increases the electron density of Ru, weakens the adsorption of H2 on Ru and makes H2 not be activated, so the reaction selectivity of PROX has been improved [144]. It is worth noting that the presence of H2 in the atmosphere decreases the oxidation activity of CO on the Rh/SiO2 catalyst, but promotes the oxidation activity of CO with the addition of K promoter [164]. The same phenomenon is also found on Pt-based catalysts [162]. K reduces the adsorption of CO in the PROX process on Pt/Al2O3 catalyst, the adsorption species of H2 and O2 (such as OH species) are formed on the surface, and the OH species participate in the oxidation of CO, so that the presence of H2 in the atmosphere promotes the oxidation of CO [142].

3.2.3. Influencing Factors on Promotion Effect

The promotion effect of K is affected by the type and content of alkali metal, preparation method and carrier. The interaction between surface OH groups and alkali metal ions largely depends on the alkalinity, while K performs best promotion effect on Pt/Al2O3 due to its moderately alkaline [162].
The content of K also affects the promotion effect. The volcanic efficiency is well reflected in the content and the promotion effect on CO catalytic oxidation catalyst [165]. On Pt/TiO2, the promotion effect is not obvious when K content is 0.1 wt%, but it is obviously promoted when K content is 0.3 wt%, while the activity decreases when K content increases to 0.5 wt% [26]. This is because an appropriate amount of K increases the dispersion of Pt, while an excessive amount of K reduces the dispersion of Pt. The catalyst shows the best activity as K/Rh = 3 on Rh/SiO2. At a low K content, a strong interaction between K and Rh is obtained, while, at a high K content, too much K will cover the active sites, resulting in the decrease of the number of active sites [164]. On Pt/Al2O3 catalyst, at an alkali metal content of M/Pt ≤ 3, all alkali metal ions except Li show promotion effect. At M/Pt > 3, Na and K continue to show promotion effect, but Rb and Cs show inhibitory effect [162].
The preparation method also has an effect on the promotion effect. The K-modified Ru/SiO2 catalysts prepared by co-impregnation, sequential impregnation and sequential calcination are compared, and the catalysts prepared by sequential impregnation show the best activity [164]. A suitable preparation method is suggested to improve the dispersion of Rh and the interaction between Rh and K2CO3 [164].
For the same active center metal, K performs different promotion effect on different carriers. For Rh-based catalysts, K promotes the PROX activity on Rh/SiO2 and Rh/USY catalysts, but the promotion mechanism is different. On the Rh/SiO2 catalyst, K improves the dispersion of Rh with turnover frequency (TOF) not change, while on the Rh/USY catalyst, K improves the TOF with the dispersion of Rh unchanged [154]. Compared with the CuO-CeO2 catalysts supported on CNT and reduced graphene oxide, it is found that K plays a better promotion effect on CNT, which is attributed to the fact that CNT has a better electron transfer effect than reduced graphene oxide [160].
In the two similar reactions of soot oxidation and CO oxidation, the promotion effect of K is consistent to a certain extent. K promotes the adsorption and dissociation of O2, and changes the reaction pathway. In addition, K increases the content of active center, improves the reducibility of active center, promotes the formation of active oxygen species, and improves the selectivity of CO2 and the stability of the catalyst for the two kinds of oxidation catalysts. This provides a meaningful guidance for the application of K in oxidation catalysts. However, there is a lack of research on the state of K on the catalyst surface in CO oxidation. In different catalysts, especially different active centers, K shows opposite effects on the adsorption of CO and O2. This reflects that different forms of K on the catalyst surface play different roles. More in-depth characterization and analysis of the existing state of K should be carried out.
Table 7. Summary of catalysts and reaction conditions for CO oxidation catalysts promoted by alkali metals.
Table 7. Summary of catalysts and reaction conditions for CO oxidation catalysts promoted by alkali metals.
CatalystsPromoterContent (wt%)Reaction ConditionsReferences
Temperature Range (°C)T50 a (°C)CO (%)O2 (%)H2 (%)OthersGHSV b or WHSV c
Pt/Al2O3K0.02–0.42120–28016511\\9600 h−1[140]
Pt/Al2O3K10 d50–200800.20.275\30,000 h−1[142]
Pt/Al2O3K10–20 d100–200<1000.450.4550.250.34% CH4, 16.55% CO212,000 h−1[147]
Pt/Al2O3K1–20 d100<1000.20.275\30,000 h−1[162]
Pt/Al2O3K5–20 d50–180700.20.275\30,000 h−1[153]
Rh/SiO2
Rh/USY
K350–1501000.20.275\0.015 g s mL−1[154]
Rh/USYK1–10 d80–1501000.20.275\0.015 g s mL−1[157]
Au/Al2O3Li, Rb1–30 e50–300<502.671.33\\2500 h−1[156]
Rh/SiO2K1–10 d50–2001200.20.275\0.015 g s mL−1[164]
CuO-CeO2/CNT
CuO-CeO2/graphene
K0.5–250–2001101150\0.03 g s mL−1[160]
Ru/SiO2K0.17–0.71 d80–220<801150\0.15 g s mL−1[144]
a T50 represents the temperature when the efficiency is 50%. b GHSV means gaseous hourly space velocity (h−1). c WHSV means weight hourly space velocity (g s mL−1). d The molar ratio of alkali metal to active center. e The atomic ratio of alkali metal to Al.

3.3. WGS

Water-gas shift reaction is the reaction of CO with H2O to form CO2 and H2, as shown in Equation (11). This is an important reaction to produce and purify H2. The commonly used catalysts are Cu-based, Pt-based and Co-based catalysts. The reaction conditions are shown in Table 8. The WGS reaction is divided into three processes, the adsorption of CO and H2O, the reaction of CO with H2O on the catalyst to form CO2 and H2, and the desorption of CO2 and H2.
C O + H 2 O C O 2 + H 2 .

3.3.1. Promotion Effects of K on Reaction Process

The promotion effect of K on WGS reaction has been widely found. K promotes the adsorption and dissociation of reactants, the production and decomposition of intermediates and the selectivity of CO2 products. K promotes the adsorption of CO gas molecules by adsorbing CO to bridge CO32- on Ni/CeO2 [18]. On Mo2C catalysts, the results of high resolution electron energy loss spectroscopy (HREELS), auger Electron Spectroscopy (AES) and TPD show that K stabilizes and activates the adsorbed CO, so that most of the CO is dissociated rather than desorbed among 77–277 °C [166]. K also promotes the adsorption of CO through the electrostatic interaction between Kδ+ and Oδ- [27]. Another effect is altering the adsorption of reactants. When K2CO3 is used to modify the Ru/C catalyst, a large amount of water is adsorbed around the active site, which reduces the strong adsorption of CO and balances the adsorption of H2O and CO on the active site, thus increasing the WGS activity [167].
K promotes the dissociation of H2O and decomposition of intermediates [14]. The H182O-adsorbed TPD characterization shows that the addition of K to Co-based catalysts obviously promotes the dissociation of H2O, thus increasing the activity of WGS [168]. The promotion on the dissociation of H2O, the limited process of WGS, can reduce the apparent activation energy and increase the catalytic activity [27]. In addition, K favors the thermochemistry for water dissociation on Cu/TiO2 with the cleavage of an O-H bond occurring at room temperature [169]. The addition of K on Pt-based catalysts weakens the strength of C-H bond in intermediate formate and promotes the decomposition of intermediate products [170].
K also improves the selectivity in the reaction process, and the mechanism of improving the selectivity on Ni-based catalysts is shown in Figure 14 [171]. On the non-K catalyst, methanation reaction occurs between CO2 adsorbed on the carrier and H adsorbed on Ni; on the K-modified catalyst, the K locates near Ni, adsorbs OH groups, and then inhibits the methanation reaction. The Pt-based catalysts modified by KOH has also been found to improve the selectivity of CO2 [172].

3.3.2. Promotion Effects of K on Catalyst Properties

K has a significant effect on the properties of WGS catalyst, such as the stability, type, and dispersion of active sites, as well as surface reactive oxygen species. The addition of K promoter to cobalt carbide catalysts maintains the carbonized phase of Co and stabilizes the active center [14]. In addition, the promotion effect of K lies in the formation of active centers. The key steps of WGS on Pt-based catalysts are CO adsorption and H2O activation, both of which occur at the oxidized Pt site. New active species with higher activity at low temperature can be formed by adding K, especially on Pt-based and Au-based catalysts. And the new active sites can be summarized as Pt-alkali-Ox(OH)y [173] and AuOy(OH)z(K)x [90]. Based on the results of density functional theory (DFT), PtK6O4(OH)2 is considered to be the most promising candidate [173].
There are many sources of the promotion effect of K on the surface active center. After the addition of K to cobalt molybdenum carbide catalysts, the activity of WGS reaction has been increased, in that the electron donor of K promotes the electronic properties of active sites [168]. K stabilizes the dispersion of Pt atoms on Pt/SiO2 and Pt/CeO2 [174,175]. K increases the reducible oxygen on the surface of Pt/SiO2 and Pt/Al2O3, thus improving the activity [173]. This is similar to that K weakens the bond between oxygen and the surface of Al2O3 [175]. In addition to the active center, the promotion effect of K on the carrier is also proposed. K modifies the properties of Al2O3 and TiO2 on Pt/Al2O3 and Pt/TiO2 resulting increasing the catalytic activity, but remaining the rate per mole of surface Pt and the WGS kinetics [176].
K also improves the stability of the catalyst for WGS reaction. Comparing the changes of C/(Co + Mo) and O/(Co + Mo) before and after the reaction, it is found that the addition of K on cobalt molybdenum carbide catalysts significantly reduces carbon deposition and promotes the formation of CO2 [168].

3.3.3. Influencing Factors on Promotion Effect

The promotion effect of K on WGS reaction catalyst is affected by the type and content of alkali metal, precursor, and temperature. On Co-Mo/Al2O3 catalyst, K is one of the best promoters for WGS reaction, and for the precursor of K, K2CO3 is found to be the most stable source [177]. The promotion effect of alkali metal species on Co3O4 catalyst in WGS reaction follows the order of K > Na > Rb > Cs, and K also shows the best promotion effect [14,178]. However, the order of alkali metal promotion effect is different on Cu(111) and Cu(110), following the order of Cs > Rb > K > Na. And the promotion effect of K is related to work function and surface dipole moment [27]. The lower work function and the higher surface dipole moment can lead to an obvious promotion effect.
There is a volcanic relationship between promotion effect and the content. Adding 0.20–5.89 wt% of K to the Co3O4 catalyst, the promotion effect gradually increases with the increase of K content. The best promotion effect is obtained when the K content is 3.93 wt%, and then the promotion effect gradually decreases with the increase of K content [178]. The same trend is also found on Ni-based catalysts among the content of K from 1 to 10 [18]. In addition, the temperature has a great effect on the catalytic activity of the K-modified catalyst. This is due to the weak binding force of K, which is easily lost at high temperature, resulting in poor stability of the catalyst [179].
In the WGS reaction, the promotion effect of K is mainly attributed to the promoting the adsorption of CO, balancing the adsorption of CO and H2O, changing the reaction pathway, improving the selectivity of CO2, stabilizing the active center, and improving stability. The important finding is that K combines with the active center metal to form a new active center, which is rarely found in other reaction. K can reduce carbon deposition and promote the formation of CO, which provides a choice for K in the chemical industry to improve catalyst stability. There is a lack of research on the influence of precursors. Only the influence of alkali metal type is studied, but the optimal alkali metal type is different under different systems.
Table 8. Summary of catalysts and reaction conditions for WGS catalysts promoted by alkali metals.
Table 8. Summary of catalysts and reaction conditions for WGS catalysts promoted by alkali metals.
CatalystsPromoterContent (wt%)Reaction ConditionsReferences
Temperature Range (°C)T50a (°C)CO (%)H2O (%)OthersGHSV b or WHSV c
Co3O4K0.2–5.89300<3003.2216.1322.58% H20.0077 g s mL−1[178]
Co3O4Li, Na, K, Rb, Cs0.02–0.05 d180–300215326.129.9% H22.4 g s mL−1[14]
cobalt molybdenum carbideK2180<18010.52120% H20.2 g s mL−1[168]
Pt/ceriaLi, Na, K, Rb, Cs0.15–2.9225–2752502.846.750.5% H20.018 g s mL−1[170]
Pt/SiO2Na1–3150–350220210\0.09 g s mL−1[173]
Pt/Al2O3Li, Na, K7–125 d230–250\6.821.98.5% CO2, 37.4% H20.796 g s mL−1[176]
Pt/TiO2Na1–10200–300\2.835.6637.74% H20.012–0.2 g s mL−1[180]
Ni/CeO2K1–10300–6003205 mol.%25 mol.%\68,000 h−1[18]
LaNiO3K2.5–10350–550<350525\0.06 g s mL−1[171]
Ru/CK2–10200–3252951020\0.3 g s mL−1[167]
a T50 represents the temperature when the efficiency is 50%. b GHSV means gaseous hourly space velocity (h−1). c WHSV means weight hourly space velocity (g s mL−1). d The molar ratio of alkali metal to active center.

3.4. RWGS

Reverse water-gas shift reaction is the reverse reaction of WGS, used to convert CO2 into CO, which is an important raw material for chemical production. The reaction is shown in Equation (12). The commonly used catalysts are Fe-based, Pt-based, and Ni-based catalysts, and the reaction conditions are shown in Table 9.
C O 2 + H 2 C O + H 2 O .

3.4.1. Promotion Effects of K on Reaction Process

K shows a promotion effect on RWGS reaction [181], which is embodied in the adsorption of reactants, the change of reaction pathway and the selectivity of product CO. The alkalinity of K promotes the adsorption of acidic reactants CO2 [182,183]. As an electron promoter, K promotes electron transfer and enhances the electrostatic interaction between the catalyst and reaction molecules [184]. It is found that the stable intermediates additionally appear after adding K with the reaction pathway changed on Fe/Al2O3 [185]. For the Fe-based catalysts, RWGS follows the redox mechanism before modification, as shown in Equations (13)–(17), and an associative pathway completes the reaction after adding K, as shown in Equations (18)–(22) [185]. The Δ in equation represents the active site. On Pt-based catalysts, K weakens the adsorption strength of CO on Pt through electronic properties, resulting in a weak interaction between CO and Pt which promotes the formation of formic acid intermediates [186]. K improves the selectivity of CO in the reaction process [187]. It is found that K increases the binding energy, improves the reaction activity of CO2, and promotes the cleavage of C = O bond and the formation of CO on Mo2C [188].
C O 2 g + Δ C O 2 Δ ,
C O 2 Δ + Δ C O Δ + O Δ ,
C O Δ C O g + Δ ,
H 2 g + O Δ H 2 O Δ ,
H 2 O Δ H 2 O g + Δ ,
H 2 g + Δ H 2 Δ ,
H 2 Δ + Δ H Δ + H Δ ,
C O 2 g + Δ C O 2 Δ ,
C O 2 Δ + H Δ C O O H Δ + Δ ,
C O O H Δ + H Δ C O g + H 2 O g + 2 Δ .

3.4.2. Promotion Effects of K on Catalyst Properties

K has an important influence on the dispersion, stability and activity of the active center. K improves the dispersion of Ni on Ni/Al2O3 [181]. And K can keep the active center Mo in the reduced state and active state on Mo-based catalysts [182]. K forms new active sites at the interface of Cu-K on Cu/SiO2. Moreover, the addition of K can form K2O, which is considered to be the active site for the decomposition of formates [189]. H2-TPR and XPS results show that a strong interaction between K and Pt stabilizes Pt in a high oxidation state on Pt-based catalysts [186]. The new interface between KOx and Pt formed by adding K is considered as the active site for the decomposition of formic acid to produce CO [186]. Consistent with the WGS reaction, XANES and XPS results show that the addition of K promotes the formation of Pt-O(OH)-K, and in-situ DRIFT spectroscopy and microcalorimetry measurements determine that the Pt-O(OH)-K interface is the main active site to adsorb CO2 and produce bridge-bonded formate intermediates [190].

3.4.3. Influencing Factors on Promotion Effect

The promotion effect of K on RWGS reaction is affected by the content of K. The atomic ratio of K/Pt from 5 to 200 is studied on Pt-based catalysts, and the promotion effect is the best when the atomic ratio of K/Pt is 80, and too much K will have a blocking effect, thereby weakening the promotion effect [190].
K shows promotion effect on the two reverse reactions of WGS and RWGS, but there are similarities, as well as differences, between the two kinds of reactions. The similar influence of K is forming active centers. Taking the Pt-based catalysts as an example, K forms the Pt-O(OH)-K active species and promotes the formation of the reaction intermediate formate species. Interestingly, the formed Pt-O(OH)-K species can promote both WGS and RWGS. The difference lies in the catalytic temperature, as shown in Table 8 and Table 9. The promotion of K for WGS is shown at low temperature (<400 °C), while, for RWGS at high temperature, (>400 °C). In addition, K shows the ability to weaken the C-H bond in WGS reaction, while shows the ability to weaken the C = O bond in RWGS reaction. The promotion effect of K for WGS and RWGS both originates in the formation of intermediate formate species. However, there is a lack of research on the influence of type of alkali metals and precursors.
Table 9. Summary of catalysts and reaction conditions for reverse water-gas shift (RWGS) catalysts promoted by alkali metals.
Table 9. Summary of catalysts and reaction conditions for reverse water-gas shift (RWGS) catalysts promoted by alkali metals.
CatalystsPromoterContent (wt%)Reaction ConditionsReferences
Temperature Range (°C)T20a (°C)CO2 (%)H2 (%)GHSV b or WHSV c
Fe/Al2O3K1–4480\15/6060/150.018–0.036 g s mL−1[185]
Fe/Al2O3Cs0–5400–800<400CO2:H2 = 1:40.3 g s mL−1[191]
Ni/Al2O3K2400–70040050500.24 g s mL−1[181]
Pt/zeoliteK5–200 d200–50045045450.12 g s mL−1[190]
Mo2C/γ-Al2O3K1–3300<300CO2:H2 = 1:30.24–1.2 g s mL−1[182]
WC/γ-Al2O3Na, K0.25d350350CO2:H2 = 1:30.4–4 g s mL−1[187]
Co-CeO2K1400–600425CO2:H2 = 1:10.012 g s mL−1[192]
Cu/SiO2K0.52–5.2200–600>6005050\[189]
a T20 represents the temperature when the efficiency is 20%. b GHSV means gaseous hourly space velocity (h−1). c WHSV means weight hourly space velocity (g s mL−1). d The molar ratio of alkali metal to active center.

4. Removal and Reuse of VOCs

VOCs produced by industrial production has a serious impact on the environment and human body. Some of them are toxic and carcinogenic. VOCs is also a contributing factor to haze and photochemical smog. The methods of removal and reuse of VOCs include catalytic oxidation and reforming reaction.

4.1. VOCs Oxidation

VOCs catalytic oxidation is that VOCs are oxidized to small molecules, such as CO2 and H2O, and may simultaneously produce by-products, such as CH4 and CO. Taking formaldehyde as an example, the reaction process is shown in Equation (23). Catalytic oxidation is commonly used to remove VOCs in motor vehicle exhaust and industrial flue gas. And partial oxidation of VOCs is the important way to produce chemical raw materials. The commonly used catalysts are Pt-based, Cu-based, Ni-based, and Co-based catalysts, and the reaction conditions are shown in Table 10.
H C H O + O 2 H 2 O + C O 2 .

4.1.1. Promotion Effects of K on Reaction Process

The promotion effect of K on the catalytic oxidation of VOCs has been widely studied [193,194,195]. These includes the promotion of the affinity of oxygen species, the decomposition of reaction intermediates, the influence of reaction pathway, and the promotion of product selectivity. The in-situ DRIFT results show that the addition of K enhances the decomposition of intermediate formate on Pt/Al2O3 catalysts [196]. The addition of K to Pt-based catalysts promotes the decomposition of intermediate formaldehyde in the dichloromethane oxidation reaction [170,197], in that the formation of Pt-O(OH)x-K species changes the reaction pathway of formaldehyde and the reaction equilibrium, thus accelerates the decomposition of formaldehyde to form CO [198]. For the oxidation of benzyl alcohol on Cu/NaY, K increases the type of adsorbed oxygen and provides greater affinity of oxygen [199].
K opens up a new reaction pathway for the catalytic oxidation of VOCs. In the catalytic oxidation of formaldehyde over Pt/TiO2 catalyst, K stabilizes the dispersed Pt-O(OH)x-alkali metal species on the surface, significantly improves the activity of formaldehyde oxidation by activating H2O and catalyzing the reaction of surface OH groups with formats [198]. On the non-K catalyst, the results of in-situ DRIFT show that the oxidation of HCHO follows the formate decomposition route, as shown in Equation (24). However, a new reaction pathway appears on the K-modified catalysts, the direct oxidation pathway of formate is shown in Equation (25). The reaction between surface hydroxyl group and formate is prior to the decomposition of formate into CO, and then oxidation of CO [200]. The kinetic results show that K obviously increases the reaction rate due to the new formation of OH group, which also confirms that the reaction pathway has changed [198]. The existence of OH group inhibits the direct decomposition of formate, as shown in Equation (26), and promotes the direct oxidation process of formate, as shown in Equation (27). Moreover, the existence of K continuously supplements the OH group by activating H2O. The effect of K on weakening the C-H bonds also be found on H-Zeolite Socony Mobil–5-supported Pd catalysts (Pd/H-ZSM-5) [201], similar to the finding in WGS. Similarly, the C-H bond is weakened by K observed at low temperature (<400 °C).
H C H O H C O O C O C O 2 ,
H C H O H C O O + O H C O 2 + H 2 O ,
H C O O M C O M + O H M ,
H C O O M + O H M H 2 O + C O 2 + 2 M .
K promotes the selectivity of VOCs oxidation. K improves the oxidation selectivity by reducing the reactivity of lattice oxygen on Fe-based catalysts in propylene epoxidation [202] and reduces the by-product acetaldehyde production on the Co-Mn-Al catalyst in ethanol oxidation [203]. For the propylene partial oxidation over CuOx/SiO2 catalysts, K decreases the nucleophilic strength of oxygen in CuOx by attracting its electrons, accompanied by a notable shift in selectivity from acrolein towards propylene oxide, as shown in Figure 15 [16].

4.1.2. Promotion Effects of K on Catalyst Properties

K has an effect on the active center of VOCs oxidation catalyst, such as increasing the surface active species, improving the dispersion of active center, and reducing the size of metal cluster, increasing the specific surface area, improving the basicity and redox properties, and improving the stability of the catalyst. K can form a new species Pt-O-Kx on the surface to enhance the reducibility of the catalyst on Pt/Al2O3 catalyst in dichloromethane oxidation [173]. It is further found that the washing process has no effect on the activity, indicating that K plays a role by connecting with the catalyst, which indirectly proves the formation of new species Pt-O-Kx [198].
K has an effect on the properties of the active center of VOCs oxidation catalyst. The addition of KCl changes the local coordination of Fe through electronic interaction on FeOx/SBA-15, resulting in the formation of tetrahedral configuration of Fe [202]. K also improves the dispersion of active center on Fe-based catalysts [202], Au-based catalysts [204], and Pd-based catalysts [201]. And the EXAFS results show that K reduces the particle size of Pt on Pt/TiO2 [198]. K also increases the specific surface area and basicity on NiCo2O4 catalyst [205] and Co-Mn-Al catalyst [203].
The electron supply of K improves the reactivity of O in M-O bond, so the catalytic oxidation activity can be promoted [137]. Moreover, the effect of K on the increase of valence state of active center has been widely found on Co-Mn-Al [203], Co/NaY [206], Co/NaUSY [207], and Pt/Al2O3 [196]. And K can increase the oxidation performance by increasing the electrophilic oxygen species [205] and forming hydroxyl species [173,180] on the surface. In addition, K also improves the stability of the catalyst. Compared with catalysts without K, the hydroxyl species appear near Pt stabilized by K. The addition of K to the Pt/TiO2 catalyst used in the oxidation reaction of formaldehyde improves the stability in humid atmosphere [198].

4.1.3. Influencing Factors on Promotion Effect

The effect of K on the catalytic oxidation of VOCs is affected by the type, precursor, the content and atmosphere. The K performs the best promotion effect on FeOx/SBA-15 than Na, Rb, and Cs [202]. And KCl is the suitable precursor than K2CO3, KNO3 and KBr. The content corresponding to the best effect is different in different types of VOCs oxidation. The effect of 0–3 wt% content K on Co-Mn-Al catalyst is studied. The K content of 1 wt% shows the best activity in toluene oxidation, while the activity increases with the increase of K content in ethanol oxidation [203]. The volcanic relationship between promotion effect and the content is found on Pt-based catalysts among the ratio of K/Al from 0.02 to 0.1 [208]. The inhibition of excessive K is attributed to the hindrance of the formation of a large number of three-dimensional crystallites on the surface [209].
The ratio of reactants to O2 in the atmosphere affects the promotion effect of K. On the Pt/Y2O3-ZrO2 catalyst, under the condition of C3H6/O2 near to the stoichiometric ratio, K shows a strong promotion effect [210]. However, when O2 is in large excess, K loses its promotion effect, or even performs an inhibitory effect [211], while K shows a strong promotion effect on Pt/β-Al2O3 catalyst under stoichiometric ratio and oxygen-rich conditions [212].
The promotion effect of K on VOCs oxidation is similar to that of soot oxidation and CO oxidation. Among them, K affects the reaction pathway, product selectivity, redox properties of catalysts, and the formation of active oxygen species. Different from soot oxidation and CO oxidation, the addition of K to catalysts for the VOCs oxidation plays a role in promoting the dissociation of intermediate species, thus promoting the activity of the reaction. The weakening of C-H bond is also found at low temperature (< 400 °C). In VOCs oxidation, new species of K formed on Pt-based catalysts are also commonly found. However, it is found that K can increase the metal valence in VOCs oxidation, which is inconsistent with the results in the N2O and NO decomposition, and further research is needed. Due to the different target reactants of VOC oxidation, the promotion effect of K content shows inconsistency. It is worth noting that the application of K-modified catalyst to VOCs oxidation in industrial flue gas needs to consider the effect of atmosphere on K promoter in detail.
Table 10. Summary of catalysts and reaction conditions for volatile organic compounds (VOCs) oxidation catalysts promoted by alkali metals.
Table 10. Summary of catalysts and reaction conditions for volatile organic compounds (VOCs) oxidation catalysts promoted by alkali metals.
CatalystsPromoterContent (wt%)Reaction ConditionsReferences
Temperature Range (°C)T50a (°C)VOCsO2 (%)OthersGHSV b or
WHSV c
Pt/Al2O3K0.02–0.05225–5002753000 ppm Dichloromethane\air15,000 h−1[196]
Pt/TiO2Na1–215–200<15600 ppm Formaldehyde2050% Relative humidity120,000 h−1[198]
Pt/Al2O3K0.02–0.1f100–350200500 ppm ethanol\air0.072 g s mL−1[208]
Co/NaY
Co/SiO2
Li, Na, K, Rb, Cs2–16350>3502.78% benzyl alcohol8.33N2520 g s mol−1[207]
Co/NaY
Co/NaUSY
K0–25 d350>3502.78% benzyl alcohol8.33N2520 g s mol−1[206]
Co-Mn-AlK0–3100–4001451 g/m3 toluene and ethanol\air0.36 g s mL−1[203]
Co0.1/ZrO2Cs0.15\2961000 ppm toluene10N210,500 h−1e[213]
NiCo2O4K2220–400330(T98)Toluene; acetone; alcohol; acetic ether\air5000 h−1[205]
Cu-Mg-AlK0.9250–6004150.5% Methane4.5He\[214]
Cu/ZrO2Cs0.15250–5004801000 ppm toluene10N210,500 h−1e[215]
Fe/SBA-15Li, Na, K, Rb, Cs5d320>3202.5% Propylene\25% N2O0.2 g s mL−1[202]
Natural manganese oreK0.07200–330225550ppm o-xylene20%N20.6 g s mL−1[216]
a T50 represents the temperature when the efficiency is 50%. b GHSV means gaseous hourly space velocity (h−1). c WHSV means weight hourly space velocity (g s mL−1). d The molar ratio of alkali metal to active center. e This represents a liquid hourly space velocity (LHSV). f The represents the ratio of K and Al.

4.2. Reforming Reaction

The reforming reaction is the reorganization of organic molecules into gaseous small molecules, such as CO, H2, etc. The steam reforming and CO2 reforming are shown in Equations (28) and (29). The reforming reaction is the important way to produce chemical raw materials. The commonly used catalysts are Pt-based catalysts and Ni-based catalysts. The reaction conditions are shown in Table 11.
C x H y O z + H 2 O C x H y O z + H 2 + C O + C O 2 ,
C x H y O z + C O 2 C x H y O z + H 2 + C O .

4.2.1. Promotion Effects of K on Reaction Process

K plays an important role in the reforming reaction, such as the adsorption of reactants, the selectivity of products, and reaction pathway. The addition of K can promote the chemisorption of CO on the surface of Mo carbide and Mo(100) [166]. Because of its alkalinity, K also promotes the adsorption of CO2 [217]. The addition of K can promote the activity and stability of the reforming of methanol, and K is a good promoter to increase the production of CO2 on the Ni/Al catalyst [19]. The promotion effect of K on selectivity is also found on Mo2C [218] and Pt/Al2O3 [219]. The acidic sites of alumina catalyze the dehydration of ethanol, and the addition of K neutralizes these acidic sites thus inhibits the formation of ethylene, which is beneficial to improve the selectivity of the reforming of methanol [208]. Through TPD, it is found that K reduces the stability of surface acetate species, so that they can be decomposed at lower temperatures to produce the CO2 and CH4, favoring to improve the selectivity [219]. K changes the ethanol adsorption from bridged form acetate to monodentate form acetate on the Co/α-Al2O3, thereby increasing selectivity of ethanol to H2 [220]. The surface acetate can inhibit the steam reforming activity of Pt-based catalysts and lead to a significant decrease in the number of products, but also promote the formation of ethylene. The promotion effect of K on the decomposition of acetate significantly enhances the selectivity of the target products (H2 and CH4) of steam recombination [219]. On K-promoted Pt-hydrotalcite catalysts (K-Pt-HT) for reforming of glycerol to the production of H2, K changes the reaction pathway and reduces the formation of by-products, thereby increasing the selectivity of H2, as shown in Figure 16 [221].

4.2.2. Promotion Effects of K on Catalyst Properties

More and more attention has been paid to the effect of K on the active center of the reforming catalyst. It is embodied in the generation of metal-carrier interaction, the change of redox, the increase of the content of active sites and the improvement of stability. It has been discovered that K leads to strong metal to support interaction (SMSI), especially on Ni/Al2O3 [222,223]. The H2-TPR results show that K shifts the reduction peak of the catalyst to high temperature, which means that the interaction between the active center and the carrier becomes stronger. And adding K can increase the number of active Ni sites and realize a higher activity on the Ni/Al2O3 catalyst [224].
K also significantly improves the stability of the catalyst for the reforming reaction. The role of K in preventing the occurrence of coking on the catalyst surface is widely recognized [225]. It can be attributed to the ability of K to divide the surface into small parts [226], promote CO2 adsorption and strengthen carbon gasification [217,224], strengthen the boudouard reaction [227] which C reacts with CO2 to produce CO, reduce acid sites on the carrier surface [228] and depress the metal sintering [229].

4.2.3. Influencing Factors on Promotion Effect

The promotion effect of K on the catalysts for reforming reaction is affected by the type and the content of K. It is found that K shows the best promotion effect than other alkali metals on Ni-based catalysts [230]. Moreover, K shows a low coke deposition and a high catalytic activity on Ni/γ-Al2O3 catalyst [223]. The addition of K with 1, 5, 10 wt% content to the Ni/Al2O3 catalysts for methane reforming shows that the K content with 1 wt% and 5 wt% shows promotion on activity, while the K content with 10 wt% shows inhibition on activity [226].
The role of K in the reforming reaction is similar to that in WGS and RWGS reactions. All of them can promote the adsorption of reactants and the formation of reaction intermediates, and improve the selectivity of products. In the removal and reuse of VOCs, the promotion effect of K has been well found in the formation of new active centers, the decomposition of intermediates, and the generation of new reaction pathways. Due to the wide variety of VOCs, the promotion effect of K on the change of reaction pathway will be changeable. DFT is an advantageous method to clarify the specific role of K in reaction pathway and the change of the existing state of K. In the reforming reaction, the discovery that K leads to SMSI is noteworthy, especially on Ni-based catalysts. In addition, there is lack of research on the influence of the precursors. The wide variety of VOCs makes the research on the type of alkali metals and precursors become more important.
Table 11. Summary of catalysts and reaction conditions for reforming catalysts promoted by alkali metals.
Table 11. Summary of catalysts and reaction conditions for reforming catalysts promoted by alkali metals.
CatalystsPromoterContent (wt%)Reaction ConditionsReferences
Temperature range (°C)T90 a (°C)VOCsCO2 (%)H2O (%)GHSV b or WHSV c
Ni/Al2O3K0.04–0.69 d700>70050% Methane50\22,500 h−1[223]
Ni/Al2O3Na, K0–18350–6004006–7.5 f Acetic acid, ethanol, 1-propanol, propanoic acid12.1 h−1 g[224]
Ni/Al2O3K0.5–2.9800<800Methane: CO2 = 1:11.26 g s mL−1[231]
Ni/MgOLi, Na, K1650<6508% Ethanol\840,000 h−1[232]
Ni-La/cordieriteK5 e500–7006503.5 f kerosene2300 h−1[228]
Ni incorporated mesoporous smectiteLi, Na, K, Rb, Cs1450<4503.3 f Acetic acid\270.005 g s mL−1[230]
Pt/Al2O3K0.04–0.4450<450Ethanol : H2O = 1:30.05 g s mL−1[219]
a T90 represents the temperature when the efficiency is 90%. b GHSV means gaseous hourly space velocity (h−1). c WHSV means weight hourly space velocity (g s mL−1). d The molar ratio of alkali metal to active center. e The weight content percentage (wt%) of K2O. f This represents a given steam to carbon ratio. g This represents a liquid hourly space velocity (LHSV).

5. Summary

To summarize, K shows promotion effect in many kinds of reactions, and plays an important role in the removal and reuse of NOx, COx (x = 0, 1, 2), and VOCs. The promotion effect of K on the catalytic reaction is mainly reflected in three aspects: the reaction process, structural properties and chemical properties of catalysts. The promotion effects on the reaction process include the adsorption of reactants, desorption of products, reaction pathway, recovery of active center, and selectivity. The promotion effects on the catalyst properties involve improving basicity, increasing the content of active centers, improving stability, improving the electron donor characteristics, changing the redox property, increasing the surface area and promoting SMSI. The bibliometrics is an effective method to analyze and reflect the relationship between different factors [233]. The correlation results of the promotion effect of K on the commonly used active centers and types of promotion effect for different pollutants are shown in Figure 17. It can be clearly seen that, on Co-based, Mn-based, and Pt-based catalysts, K shows the promotion effect of the removal and reuse of the all types of pollutants. In addition, on Fe-based, Cu-based, Mo-based, and Ni-based catalysts, there are more research on the promotion effect of K in the removal and reuse of COx and VOCs. For the types of metals, the research of transition metals is significantly more than that of alkaline earth metals and noble metals. The correlation between the types of promotion effect and pollutants can reflect the importance of specific promotion effect for a certain pollutant. The promotion effects on reactant, pathway, stability and active center are important for all three pollutants. Additionally, the promotion effects on electronic properties are important for NOx, surface area, and basicity are important for COx, and basicity and redox properties are important for VOCs.
The promotion effect of K is affected by many factors. These factors in different reactions are summarized in Table 12. All investigations focus on the effects of content, some of the investigations involve in the types of alkali metals and precursors. However, the effects of temperature, calcination time, preparation, carrier, and atmosphere, such as H2O, SO2, and As, have not been studied enough. According to current studies, K exhibits an excellent promotion effect and precursor K2CO3 has a better effect. The best promotion effect can be obtained by controlling the load content and calcination temperature reasonably. In most cases, the promotion effect and the content show a volcanic relationship, and lower temperature favors the stability and promotion effect of K.
The effect of K on CO adsorption and metal valence is determined by conditions. The addition of K to Rh-based, Au-based, Ni-based, and Mo-based catalysts promotes the adsorption of CO which accepts electrons on the catalyst surface, attributed to the electron-donating properties. However, the addition of K to Pt-based and Ru-based catalysts weakens the strong adsorption of CO, and relatively increases the adsorption of O2 and H2O and the cover of surface -OH species, through the electronic modification of the active center, accompanied by the weakening the metal-C bond and promoting the desorption of CO, particularly for Pt-based catalysts. However, it has also been found that the Pt-O-Kx species formed on the Pt-based catalyst can enhance the adsorption strength of CO, which reduces the energy barrier of CO oxidation and increase the reaction rate. K shows an excellent effect in regulating the active center of the catalyst. On the one hand, K supplies electrons to the metal, increases the electron density of the metal, weakens the metal-O bond, and makes the metal transfer to the low valence state. On the other hand, K increases the affinity of active sites for oxygen [207] and promotes the formation of high valence metals. The effect of K on adsorption of CO and valence state of active center needs to be studied in different catalytic systems and atmospheres. The different performances of promotion effect of K should be clearly considered in the application of industrial flue gas.

6. Potential

Although K shows excellent promotion effects in many reactions, some disadvantages need to be paid special attention in the industrial application. K volatiles above 500 °C, so high temperature windows should be avoided. In addition, K also exhibits poisoning effects, especially on V-based catalysts in SCR. Therefore, it is also necessary to consider the negative effects of K on the catalyst. In addition, most of the researches on K promoters focus on finding the relationship between the catalyst properties and activity. There are relatively few studies on the effects of K in the intermediates and the existing states of K. Clarification of the role of K in the catalytic process can better provide theoretical basis and guidance for the use of K.
The promotion effect of K on various reactions provides an idea for the synergistic removal of multiple pollutants. Most of researches are based on the promotion effect of K for a single pollutant removal reaction, but there are many gas components coexist in the industrial flue gas which perform the oxidation properties, such as O2 and VOCs, or the reduction properties, such as NO and CO. Injecting NH3 into the flue gas provides atmospheres for NH3-SCR, CO oxidation, VOCs oxidation, and NO-CO reaction. K has a potential promotion effect on the above reactions. In the aspect of synergistic removal of CO, NO, and VOCs, the addition of K can improve the removal efficiency of multi-pollutants. For the CO oxidation, SCR and VOCs oxidation, the three types of reactions have the same temperature range of 100–200 °C, so it is possible for K to promote catalysts to synergistically remove multi-pollutants. The temperature range also ensures the application of carbon-based catalysts. And activated carbon itself also has the ability to remove multi-pollutants [234,235,236], such as SO2, NOx, and VOCs. The promotion effect of K provides support for the coupling control of multi-pollutants in industrial flue gas, and provides a new idea for more economical and convenient removal of industrial flue gas pollutants.

7. Conclusions

By summarizing the promotion effect of K on the removal and reuse of NOx, COx, and VOCs, it is found that K has an important influence on reactants, intermediates and catalyst active centers based on its two important properties, basicity and electronic properties. Thus, it has a positive effect on the reaction activity, selectivity and stability of the catalyst. The effect on the reactants is mainly to promote the adsorption, which is achieved through alkalinity, increasing the adsorption sites, changing the adsorption type, and generating electrostatic interaction between Kδ+ and Oδ-. The effect on intermediates is achieved by promoting the formation of intermediates, accelerating the decomposition of intermediates and changing the type of intermediates. The effect on the active center is achieved by changing the electronic and acid-base characteristics, increasing the number of active centers, improving the dispersion, changing the valence state, forming new active centers, and recovering the active centers. The existence and combination of K are important information for promotion effect, and it is worthy of more in-depth research by DFT and EXAFS. K has a better promotion effect than other alkali metals in most cases. However, the existence of K has positive and negative effects, along with the promotion effect and content of K showing a volcanic relationship in most reactions. The application of K in industrial flue gas conditions also requires in-depth consideration of the influence of the atmosphere, which is lacking in the current research, especially for the resistance to H2O, SO2, and As. Nevertheless, K still has good prospects for applications in the removal of multiple pollutants from industrial flue gas, especially CO, NOx, and VOCs.

Funding

This research was funded by the National Key Research and Development Program of China (grant numbers 2018YFC0213406, 2017YFC0210600) and the National Natural Science Foundation of China (grant numbers U1810209).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Legutko, P.; Jakubek, T.; Kaspera, W.; Stelmachowski, P.; Sojka, Z.; Kotarba, A. Strong Enhancement of de Soot Activity of Transition Metal Oxides by Alkali Doping: Additive Effects of Potassium and Nitric Oxide. Top. Catal. 2017, 60, 162–170. [Google Scholar] [CrossRef] [Green Version]
  2. Kamal, M.S.; Razzak, S.A.; Hossain, M.M. Catalytic Oxidation of Volatile Organic Compounds (VOCs)—A Review. Atmospheric Environ. 2016, 140, 117–134. [Google Scholar] [CrossRef]
  3. Xiong, S.; Chen, J.; Huang, N.; Yang, S.; Peng, Y.; Li, J. Balance between Reducibility and N2O Adsorption Capacity for the N2O Decomposition: CuxCoy Catalysts as an Example. Environ. Sci. Technol. 2019, 53, 10379–10386. [Google Scholar] [CrossRef]
  4. Bond, T.C.; Doherty, S.J.; Fahey, D.W.; Forster, P.M.; Berntsen, T.; DeAngelo, B.J.; Flanner, M.G.; Ghan, S.; Kaercher, B.; Koch, D.; et al. Bounding the Role of Black Carbon in the Climate System: A Scientific Assessment. J. Geophys. Res. Atmos. 2013, 118, 5380–5552. [Google Scholar] [CrossRef]
  5. Liu, Y.; Shan, Y.; Wang, Y. Novel Simultaneous Removal Technology of NO and SO2 Using a Semi-Dry Microwave Activation Persulfate System. Environ. Sci. Technol. 2019, 54, 2031–2042. [Google Scholar] [CrossRef]
  6. Gálvez, M.; Ascaso, S.; Stelmachowski, P.; Legutko, P.; Kotarba, A.; Moliner, R.; Lázaro, M. Influence of the Surface Potassium Species in Fe–K/Al2O3 Catalysts on the Soot Oxidation Activity in the Presence of NOx. Appl. Catal. B Environ. 2014, 152–153, 88–98. [Google Scholar] [CrossRef] [Green Version]
  7. Lyu, X.; Guo, H.; Wang, Y.; Zhang, F.; Nie, K.; Dang, J.; Liang, Z.; Dong, S.; Zeren, Y.; Zhou, B.; et al. Hazardous Volatile Organic Compounds in Ambient Air of China. Chemosphere 2020, 246, 125731. [Google Scholar] [CrossRef]
  8. National Bureau of Statistics China. Annual Report on Environmental Statistics; National Bureau of Statistics China: Beijing, China, 2015.
  9. Sun, W.; Shao, M.; Granier, C.; Liu, Y.; Ye, C.S.; Zheng, J.Y. Long-Term Trends of Anthropogenic SO2, NOx, CO, and NMVOCs Emissions in China. Earth’s Futur. 2018, 6, 1112–1133. [Google Scholar] [CrossRef] [Green Version]
  10. Atlas, C.C. Annual CO2 Emissions. 2020. Available online: http://www.globalcarbonatlas.org/ (accessed on 1 July 2020).
  11. Mo, S.; Zhang, Q.; Sun, Y.; Zhang, M.; Li, J.; Ren, Q.; Fu, M.; Wu, J.; Chen, L.; Ye, D. Gaseous CO and toluene co-oxidation over monolithic core–shell Co3O4-based hetero-structured catalysts. J. Mater. Chem. A 2019, 7, 16197–16210. [Google Scholar] [CrossRef]
  12. Neha; Prasad, R.; Singh, S.V. A review on catalytic oxidation of soot emitted from diesel fuelled engines. J. Environ. Chem. Eng. 2020, 8, 324–331. [Google Scholar] [CrossRef]
  13. Dongil, A.; Bachiller-Baeza, B.; Castillejos, E.; Escalona, N.; Guerrero-Ruiz, A.; Rodríguez-Ramos, I. Promoter Effect of Alkalis on CuO/CeO2/Carbon Nanotubes Systems for the PROx Reaction. Catal. Today 2018, 301, 141–146. [Google Scholar] [CrossRef]
  14. Gnanamani, M.K.; Jacobs, G.; Shafer, W.D.; Sparks, D.E.; Hopps, S.; Thomas, G.A.; Davis, B.H. Low Temperature Water–Gas Shift Reaction Over Alkali Metal Promoted Cobalt Carbide Catalysts. Top. Catal. 2014, 57, 612–618. [Google Scholar] [CrossRef]
  15. Zhao, K.; Bkour, Q.; Hou, X.; Kang, S.W.; Park, J.C.; Norton, M.G.; Yang, J.-I.; Ha, S. Reverse Water Gas Shift Reaction over CuFe/Al2O3 Catalyst in Solid Oxide Electrolysis Cell. Chem. Eng. J. 2018, 336, 20–27. [Google Scholar] [CrossRef]
  16. Teržan, J.; Djinović, P.; Zavašnik, J.; Arčon, I.; Žerjav, G.; Spreitzer, M.; Pintar, A. Alkali and Earth Alkali Modified CuOx/SiO2 Catalysts for Propylene Partial Oxidation: What Determines the Selectivity? Appl. Catal. B Environ. 2018, 237, 214–227. [Google Scholar] [CrossRef]
  17. Sengodan, S.; Lan, R.; Humphreys, J.; Du, D.; Xu, W.; Wang, H.; Tao, S. Advances in Reforming and Partial Oxidation of Hydrocarbons for Hydrogen Production and Fuel Cell Applications. Renew. Sustain. Energy Rev. 2018, 82, 761–780. [Google Scholar] [CrossRef]
  18. Ang, M.; Oemar, U.; Kathiraser, Y.; Saw, E.T.; Lew, C.; Du, Y.; Borgna, A.; Kawi, S. High-Temperature Water–Gas Shift Reaction over Ni/xK/CeO2 Catalysts: Suppression of Methanation via Formation of Bridging Carbonyls. J. Catal. 2015, 329, 130–143. [Google Scholar] [CrossRef]
  19. Qi, C.; Amphlett, J.C.; Peppley, B.A. K (Na)-Promoted Ni, Al Layered Double Hydroxide Catalysts for the Steam Reforming of Methanol. J. Power Sources 2007, 171, 842–849. [Google Scholar] [CrossRef]
  20. Chen, C.; Cao, Y.; Liu, S.; Chen, J.; Jia, W. SCR Catalyst Doped with Copper for Synergistic Removal of Slip Ammonia and Elemental Mercury. Fuel Process. Technol. 2018, 181, 268–278. [Google Scholar] [CrossRef]
  21. Liu, Y.; Wang, Y.; Xu, W.; Yang, W.; Pan, Z.; Wang, Q. Simultaneous Absorption–Oxidation of Nitric Oxide and Sulfur Dioxide Using Ammonium Persulfate Synergistically Activated by UV-Light and Heat. Chem. Eng. Res. Des. 2018, 130, 321–333. [Google Scholar] [CrossRef]
  22. Pérez-Ramírez, J.; Kapteijn, F.; Schöffel, K.; Moulijn, J. Formation and Control of N2O in Nitric Acid Production. Appl. Catal. B Environ. 2003, 44, 117–151. [Google Scholar] [CrossRef]
  23. Lykaki, M.; Papista, E.; Carabineiro, S.A.C.; Tavares, P.B.; Konsolakis, M. Optimization of N2O Decomposition Activity of CuO–CeO2 Mixed Oxides by Means of Synthesis Procedure and Alkali (Cs) Promotion. Catal. Sci. Technol. 2018, 8, 2312–2322. [Google Scholar] [CrossRef]
  24. Zhao, T.; Li, Y.; Gao, Q.; Liu, Z.; Xu, X. Potassium Promoted Y2O3-Co3O4 Catalysts for N2O Decomposition. Catal. Commun. 2020, 137, 105948. [Google Scholar] [CrossRef]
  25. Koukiou, S.; Konsolakis, M.; Lambert, R.; Yentekakis, I.V. Spectroscopic Evidence for the Mode of Action of Alkali Promoters in Pt-Catalyzed Denox Chemistry. Appl. Catal. B Environ. 2007, 76, 101–106. [Google Scholar] [CrossRef]
  26. Liu, j.; Ding, T.; Tian, Y.; Li, X. Enhanced CO Oxidation Performance over Potassium-promoted Pt/TiO2 Catalysts. Chem. J. Chin. Univ. Chin. 2018, 39, 1467–1474. [Google Scholar]
  27. Wang, Y.-X.; Wang, G.-C. A Systematic Theoretical Study of Water Gas Shift Reaction on Cu(111) and Cu(110): Potassium Effect. ACS Catal. 2019, 9, 2261–2274. [Google Scholar] [CrossRef]
  28. Zhou, H.; Huang, Z.; Sun, C.; Qin, F.; Xiong, D.; Shen, W.; Xu, H. Catalytic Decomposition of N2O Over CuxCe1−xOy Mixed Oxides. Appl. Catal. B Environ. 2012, 125, 492–498. [Google Scholar] [CrossRef]
  29. Obalová, L.; Karásková, K.; Wach, A.; Kustrowski, P.; Mamulová-Kutláková, K.; Michalik, S.; Jirátová, K. Alkali Metals as Promoters in Co–Mn–Al Mixed Oxide for N2O Decomposition. Appl. Catal. A Gen. 2013, 462–463, 227–235. [Google Scholar] [CrossRef]
  30. Zhu, Z.; Lu, G.; Yang, R. New Insights into Alkali-Catalyzed Gasification Reactions of Carbon: Comparison of N2O Reduction with Carbon over Na and K Catalysts. J. Catal. 2000, 192, 77–87. [Google Scholar] [CrossRef]
  31. Maniak, G.; Stelmachowski, P.; Zasada, F.; Piskorz, W.; Kotarba, A.; Sojka, Z. Guidelines for Optimization of Catalytic Activity of 3d Transition Metal Oxide Catalysts in N2O Decomposition by Potassium Promotion. Catal. Today 2011, 176, 369–372. [Google Scholar] [CrossRef]
  32. Inger, M.; Kowalik, P.; Saramok, M.; Wilk, M.; Stelmachowski, P.; Maniak, G.; Granger, P.; Kotarba, A.; Sojka, Z. Laboratory and Pilot Scale Synthesis, Characterization and Reactivity of Multicomponent Cobalt Spinel Catalyst for Low Temper-Ature Removal of N2O from Nitric Acid Plant Tail Gases. Catal. Today 2011, 176, 365–368. [Google Scholar] [CrossRef]
  33. Zasada, F.; Stelmachowski, P.; Maniak, G.; Paul, J.-F.; Kotarba, A.; Sojka, Z. Potassium Promotion of Cobalt Spinel Catalyst for N2O Decomposition—Accounted by Work Function Measurements and DFT Modelling. Catal. Lett. 2009, 127, 126–131. [Google Scholar] [CrossRef]
  34. Asano, K.; Ohnishi, C.; Iwamoto, S.; Shioya, Y.; Inoue, M. Potassium-doped Co3O4 Catalyst for Direct Decomposition of N2O. Appl. Catal. B Environ. 2008, 78, 242–249. [Google Scholar] [CrossRef]
  35. Maniak, G.; Stelmachowski, P.; Kotarba, A.; Sojka, Z.; Rico-Pérez, V.; Bueno-López, A. Rationales for the Selection of the Best Precursor for Potassium Doping of Cobalt Spinel Based de N2O Catalyst. Appl. Catal. B Environ. 2013, 136–137, 302–307. [Google Scholar] [CrossRef]
  36. Zabilskiy, M.; Djinović, P.; Erjavec, B.; Dražić, G.; Pintar, A. Small CuO Clusters on CeO2 Nanospheres as Active Species for Catalytic N2O Decomposition. Appl. Catal. B Environ. 2015, 163, 113–122. [Google Scholar] [CrossRef]
  37. Yakovlev, A.L.; Zhidomirov, G.M.; Van Santen, R.A. N2O Decomposition Catalysed by Transition Metal Ions. Catal. Lett. 2001, 75, 45–48. [Google Scholar] [CrossRef]
  38. Yuzaki, K.; Yarimizu, T.; Ito, S.-I.; Kunimori, K. Catalytic Decomposition of N2O Over Supported Rhodium Catalysts: High Activities of Rh/USY and Rh/Al2O3 and The Effect of Rh Precursors. Catal. Lett. 1997, 47, 173–175. [Google Scholar] [CrossRef]
  39. Buenolopez, A.; Suchbasanez, I.; De Lecea, C.S. Stabilization of Active Rh2O3 Species for Catalytic Decomposition of N2O on La-, Pr-doped CeO2. J. Catal. 2006, 244, 102–112. [Google Scholar] [CrossRef]
  40. Ohnishi, C.; Asano, K.; Iwamoto, S.; Chikama, K.; Inoue, M. Alkali-doped Co3O4 Catalysts for Direct Decomposition of N2O in the Presence of Oxygen. Catal. Today 2007, 120, 145–150. [Google Scholar] [CrossRef]
  41. Xue, L.; He, H.; Liu, C.; Zhang, C.; Zhang, B. Promotion Effects and Mechanism of Alkali Metals and Alkaline Earth Metals on Cobalt−Cerium Composite Oxide Catalysts for N2O Decomposition. Environ. Sci. Technol. 2009, 43, 890–895. [Google Scholar] [CrossRef]
  42. Dou, Z.; Zhang, H.-J.; Pan, Y.-F.; Xu, X.-F. Catalytic Decomposition of N2O Over Potassium-modified Cu-Co Spinel Oxides. J. Fuel Chem. Technol. 2014, 42, 238–245. [Google Scholar] [CrossRef]
  43. Zhang, J.; Hu, H.; Xu, J.; Wu, G.; Zeng, Z. N2O Decomposition Over K/Na-promoted Mg/Zn–Ce–cobalt Mixed Oxides Catalysts. J. Environ. Sci. 2014, 26, 1437–1443. [Google Scholar] [CrossRef]
  44. Franken, T.; Palkovits, R. Investigation of Potassium Doped Mixed Spinels CuxCo3−xO4 as Catalysts for an Efficient N2O Decomposition in Real Reaction Conditions. Appl. Catal. B Environ. 2015, 176–177, 298–305. [Google Scholar] [CrossRef]
  45. Cheng, H.; Huang, Y.; Wang, A.; Li, L.; Wang, X.; Zhang, T. N2O Decomposition over K-Promoted Co-Al Catalysts Prepared from Hydrotalcite-like Precursors. Appl. Catal. B Environ. 2009, 89, 391–397. [Google Scholar] [CrossRef]
  46. Zhang, F.; Wang, X.; Zhang, X.; Turxun, M.; Yu, H.; Zhao, J. The Catalytic Activity of NiO for N2O Decomposition Doubly Promoted by Barium and Cerium. Chem. Eng. J. 2014, 256, 365–371. [Google Scholar] [CrossRef]
  47. Wu, H.-P.; Qian, Z.-Y.; Xu, X.-L.; Xu, X.-F. N2O Decomposition over K-Promoted NiAl Mixed Oxides Derived from Hydrotalcite-like Compounds. J. Fuel Chem. Technol. 2011, 39, 115–121. [Google Scholar] [CrossRef]
  48. Scagnelli, A.; Di Valentin, C.; Pacchioni, G. Catalytic Dissociation of N2O on Pure and Ni-doped MgO Surfaces. Surf. Sci. 2006, 600, 386–394. [Google Scholar] [CrossRef]
  49. Park, P.; Kil, J.; Kung, H.; Kung, M. NO Decomposition over Sodium-Promoted Cobalt Oxide. Catal. Today 1998, 42, 51–60. [Google Scholar] [CrossRef]
  50. Karásková, K.; Obalová, L.; Kovanda, F. N2O Catalytic Decomposition and Temperature Programmed Desorption Tests on Alkali Metals Promoted Co–Mn–Al Mixed Oxide. Catal. Today 2011, 176, 208–211. [Google Scholar] [CrossRef]
  51. Obalová, L.; Karásková, K.; Jirátová, K.; Kovanda, F. Effect of potassium in calcined Co–Mn–Al layered double hydroxide on the catalytic decomposition of N2O. Appl. Catal. B Environ. 2009, 90, 132–140. [Google Scholar] [CrossRef]
  52. Haneda, M.; Kintaichi, Y.; Bion, N.; Hamada, H. Alkali Metal-Doped Cobalt Oxide Catalysts for NO Decomposition. Appl. Catal. B Environ. 2003, 46, 473–482. [Google Scholar] [CrossRef]
  53. Wu, Y.; Ni, X.; Beaurain, A.; Dujardin, C.; Granger, P. Stoichiometric and Non-stoichiometric Perovskite-Based Catalysts: Consequences on Surface Properties and on Catalytic per-Formances in the Decomposition of N2O from Nitric Acid Plants. Appl. Catal. B Environ. 2012, 125, 149–157. [Google Scholar] [CrossRef]
  54. Yuan, C.; Bin Wu, H.; Xie, Y.; Lou, X.W. (David) Mixed Transition-Metal Oxides: Design, Synthesis, and Energy-Related Applications. Angew. Chem. Int. Ed. 2014, 53, 1488–1504. [Google Scholar] [CrossRef]
  55. Zabilskiy, M.; Erjavec, B.; Djinović, P.; Pintar, A. Ordered Mesoporous CuO–CeO2 Mixed Oxides as an Effective Catalyst for N2O decomposition. Chem. Eng. J. 2014, 254, 153–162. [Google Scholar] [CrossRef]
  56. Yoshino, H.; Ohnishi, C.H.; Hosokawa, S.; Wada, K.; Inoue, M. Optimized Synthesis Method for K/Co3O4 Catalyst Towards Direct Decomposition of N2O. J. Mater. Sci. 2010, 46, 797–805. [Google Scholar] [CrossRef] [Green Version]
  57. Konsolakis, M. Recent Advances on Nitrous Oxide (N2O) Decomposition over Non-Noble-Metal Oxide Catalysts: Catalytic Performance, Mechanistic Considerations, and Surface Chemistry Aspects. ACS Catal. 2015, 5, 6397–6421. [Google Scholar] [CrossRef]
  58. Zakirov, V.; Zhang, H.-Y. A Model for the Operation of Nitrous Oxide Monopropellant. Aerosp. Sci. Technol. 2008, 12, 318–323. [Google Scholar] [CrossRef]
  59. Pasha, N.; Lingaiah, N.; Reddy, P.S.S.; Prasad, P.S.S. Direct Decomposition of N2O over Cesium-doped CuO Catalysts. Catal. Lett. 2008, 127, 101–106. [Google Scholar] [CrossRef]
  60. Polizzi, S.; Bucella, S.; Speghini, A.; Vetrone, F.; Naccache, R.; Boyer, A.J.C.; Capobianco, J.A. Nanostructured Lanthanide-Doped Lu2O3 Obtained by Propellant Synthesis. Chem. Mater. 2004, 16, 1330–1335. [Google Scholar] [CrossRef]
  61. Satsuma, A.; Maeshima, H.; Watanabe, K.; Suzuki, K.; Hattori, T. Effects of Methane and Oxygen on Decomposition of Nitrous Oxide over Metal Oxide Catalysts. Catal. Today 2000, 63, 347–353. [Google Scholar] [CrossRef]
  62. Zhu, J.; Albertsma, S.; Van Ommen, A.J.G.; Lefferts, L. Role of Surface Defects in Activation of O2 and N2O on ZrO2and Yttrium-Stabilized ZrO2. J. Phys. Chem. B 2005, 109, 9550–9555. [Google Scholar] [CrossRef] [PubMed]
  63. Jiang, M.-X.; Liu, C.-G. New Insight into the Catalytic Cycle about Epoxidation of Alkenes by N2O over a Mn–Substituted Keggin-Type Polyoxometalate. J. Mol. Graph. Model. 2017, 73, 8–17. [Google Scholar] [CrossRef] [PubMed]
  64. Zhu, Z.; Lu, G. Catalytic Conversion of N2O to N2 over Potassium Catalyst Supported on Activated Carbon. J. Catal. 1999, 187, 262–274. [Google Scholar] [CrossRef]
  65. Papista, E.; Pachatouridou, E.; Goula, M.A.; Marnellos, G.Ε.; Iliopoulou, E.; Konsolakis, M.; Yentekakis, I.V. Effect of Alkali Promoters (K) on Nitrous Oxide Abatement Over Ir/Al2O3 Catalysts. Top. Catal. 2016, 59, 1020–1027. [Google Scholar] [CrossRef]
  66. Xue, L.; Zhang, C.; He, H.; Teraoka, Y. Promotion Effect of Residual K on the Decomposition of N2O over Cobalt–cerium Mixed Oxide Catalyst. Catal. Today 2007, 126, 449–455. [Google Scholar] [CrossRef]
  67. Obalová, L.; Maniak, G.; Karásková, K.; Kovanda, F.; Kotarba, A. Electronic Nature of Potassium Promotion Effect in Co–Mn–Al Mixed Oxide on the Catalytic Decomposition of N2O. Catal. Commun. 2011, 12, 1055–1058. [Google Scholar] [CrossRef] [Green Version]
  68. Grzybek, G.; Stelmachowski, P.; Gudyka, S.; Duch, J.; Ćmil, K.; Kotarba, A.; Sojka, Z. Insights into the Twofold Role of CS Doping on De N2O Activity of Cobalt Spinel Catalyst—Towards Rational Optimization of the Precursor and Loading. Appl. Catal. B Environ. 2015, 168–169, 509–514. [Google Scholar] [CrossRef]
  69. Stelmachowski, P.; Maniak, G.; Kotarba, A.; Sojka, Z. Strong Electronic Promotion of Co3O4 Towards N2O Decomposition by Surface Alkali Dopants. Catal. Commun. 2009, 10, 1062–1065. [Google Scholar] [CrossRef]
  70. Karásková, K.; Obalová, L.; Jirátová, K.; Kovanda, F. Effect of Promoters in Co–Mn–Al mixed Oxide Catalyst on N2O Decomposition. Chem. Eng. J. 2010, 160, 480–487. [Google Scholar] [CrossRef]
  71. Grzybek, G.; Wójcik, S.; Legutko, P.; Gryboś, J.; Indyka, P.; Leszczyński, B.; Kotarba, A.; Sojka, Z. Thermal Stability and Repartition of Potassium Promoter between the Support and Active Phase in the K-Co2.6Zn0.4O4|α-Al2O3 Catalyst for N2O Decomposition: Crucial Role of Activation Temperature on Catalytic Performance. Appl. Catal. B Environ. 2017, 205, 597–604. [Google Scholar] [CrossRef]
  72. Jirátová, K.; Pacultová, K.; Balabánová, J.; Karásková, K.; Klegová, A.; Bílková, T.; Jandová, V.; Koštejn, M.; Martaus, A.; Kotarba, A.; et al. Precipitated K-Promoted Co–Mn–Al Mixed Oxides for Direct NO Decomposition: Preparation and Properties. Catalysts 2019, 9, 592. [Google Scholar] [CrossRef] [Green Version]
  73. Pacultová, K.; Bílková, T.; Klegova, A.; Karásková, K.; Fridrichová, D.; Jirátová, K.; Kiška, T.; Balabánová, J.; Koštejn, M.; Kotarba, A.; et al. Co-Mn-Al Mixed Oxides Promoted by K for Direct NO Decomposition: Effect of Preparation Parameters. Catalysts 2019, 9, 593. [Google Scholar] [CrossRef] [Green Version]
  74. Tsujimoto, S.; Masui, T.; Imanaka, N. Fundamental Aspects of Rare Earth Oxides Affecting Direct NO Decomposition Catalysis. Eur. J. Inorg. Chem. 2015, 2015, 1524–1528. [Google Scholar] [CrossRef]
  75. Hong, W.-J.; Iwamoto, S.; Hosokawa, S.; Wada, K.; Kanai, H.; Inoue, M. Effect of Mn Content on Physical Properties of CeOx–MnOy Support and BaO–CeOx–MnOy Catalysts for Direct NO Decomposition. J. Catal. 2011, 277, 208–216. [Google Scholar] [CrossRef] [Green Version]
  76. Haneda, M.; Kintaichi, Y.; Hamada, H. Reaction Mechanism of NO Decomposition over Alkali Metal-Doped Cobalt Oxide Catalysts. Appl. Catal. B Environ. 2005, 55, 169–175. [Google Scholar] [CrossRef]
  77. Haneda, M.; Nakamura, I.; Fujitani, T.; Hamada, H. Catalytic Active Site for NO Decomposition Elucidated by Surface Science and Real Catalyst. Catal. Surv. Asia 2005, 9, 207–215. [Google Scholar] [CrossRef]
  78. Jirátová, K.; Pacultová, K.; Karásková, K.; Balabánová, J.; Koštejn, M.; Obalová, L. Direct Decomposition of NO over Co-Mn-Al Mixed Oxides: Effect of Ce and/or K Promoters. Catalysts 2020, 10, 808. [Google Scholar] [CrossRef]
  79. Brown, W.A.; King, D.A. NO Chemisorption and Reactions on Metal Surfaces: A New Perspective. J. Phys. Chem. B 2000, 104, 2578–2595. [Google Scholar] [CrossRef]
  80. Pacultová, K.; Draštíková, V.; Chromčáková, Ž.; Bílková, T.; Kutláková, K.M.; Kotarba, A.; Obalová, L. On The Stability of Alkali Metal Promoters in Co Mixed Oxides during Direct NO Catalytic Decomposition. Mol. Catal. 2017, 428, 33–40. [Google Scholar] [CrossRef]
  81. Kotarba, A.; Kruk, I.; Sojka, Z. Energetics of Potassium Loss from Styrene Catalyst Model Components: Reassignment of K Storage and Release Phases. J. Catal. 2002, 211, 265–272. [Google Scholar] [CrossRef]
  82. Bieniasz, W.; Trębala, M.; Sojka, Z.; Kotarba, A. Irreversible Deactivation of Styrene Catalyst Due to Potassium Loss—Development of Antidote via Mechanism Pinning. Catal. Today 2010, 154, 224–228. [Google Scholar] [CrossRef]
  83. Peck, T.C.; Roberts, C.A.; Reddy, G.K. Contrasting Effects of Potassium Addition on M3O4 (M = Co, Fe, and Mn) Oxides during Direct NO Decomposition Catalysis. Catalysts 2020, 10, 561. [Google Scholar] [CrossRef]
  84. Karásková, K.; Pacultová, K.; Klegova, A.; Fridrichová, D.; Valášková, M.; Jirátová, K.; Stelmachowski, P.; Kotarba, A.; Obalová, L. Magnesium Effect in K/Co-Mg-Mn-Al Mixed Oxide Catalyst for Direct NO Decomposition. Catalysts 2020, 10, 931. [Google Scholar] [CrossRef]
  85. Hong, W.-J.; Iwamoto, S.; Inoue, M. Direct NO Decomposition over a Ce–Mn Mixed Oxide Modified with Alkali and Alkaline Earth Species and CO2-TPD Behavior of the Catalysts. Catal. Today 2011, 164, 489–494. [Google Scholar] [CrossRef]
  86. Xiang, J.; Du, X.; Wan, Y.; Chen, Y.; Ran, J.; Zhang, L. Alkali-driven Active Site Shift of fast SCR with NH3 on V2O5–WO3/TiO2 Catalyst via a Novel Eley–Rideal Mechanism. Catal. Sci. Technol. 2019, 9, 6085–6091. [Google Scholar] [CrossRef]
  87. Xu, Z.; Li, Y.; Guo, J.; Xiong, J.; Lin, Y.; Zhu, T. An Efficient and Sulfur Resistant K-Modified Activated Carbon for SCR Denitrification Compared with Acid-and CU-Modified Activated Carbon. Chem. Eng. J. 2020, 395, 125047. [Google Scholar] [CrossRef]
  88. Hao, Z.; Shen, Z.; Li, Y.; Wang, H.; Zheng, L.; Wang, R.; Liu, G.; Zhan, S. The Role of Alkali Metal in α-MnO 2 Catalyzed Ammonia-Selective Catalysis. Angew. Chem. Int. Ed. 2019, 58, 6351–6356. [Google Scholar] [CrossRef]
  89. Wang, X.; Wu, D.; Zhang, J.; Gao, X.; Ma, Q.; Fan, S.; Zhao, T.-S. Highly Selective Conversion of CO2 to Light Olefins via Fischer-Tropsch Synthesis over Stable Layered K–Fe–Ti Catalysts. Appl. Catal. A Gen. 2019, 573, 32–40. [Google Scholar] [CrossRef]
  90. Yang, M.; Li, S.; Wang, Y.; Herron, J.A.; Xu, Y.; Allard, L.F.; Lee, S.; Huang, J.; Mavrikakis, M.; Flytzani-Stephanopoulos, M. Catalytically Active Au-O(OH)x-Species Stabilized by Alkali Ions on Zeolites and Mesoporous Oxides. Science 2014, 346, 1498–1501. [Google Scholar] [CrossRef]
  91. Wang, C.; Wang, J.; Wang, J.; Shen, M. Promotional Effect of Ion-Exchanged K on the Low-Temperature Hydrothermal Stability of Cu/SAPO-34 And Its Synergic Application with Fe/Beta Catalysts. Front. Environ. Sci. Eng. 2020, 15, 1–13. [Google Scholar] [CrossRef]
  92. Yentekakis, I.V.; Konsolakis, M.; Lambert, R.; MacLeod, N.; Nalbantian, L. Extraordinarily Effective Promotion by Sodium in Emission Control Catalysis: NO Reduction by Propene over Na-Promoted Pt/γ-Al2O3. Appl. Catal. B Environ. 1999, 22, 123–133. [Google Scholar] [CrossRef]
  93. Konsolakis, M.; Yentekakis, I.V. Strong Promotional Effects of Li, K, Rb and Cs on the Pt-Catalysed Reduction of NO by Propene. Appl. Catal. B Environ. 2001, 29, 103–113. [Google Scholar] [CrossRef] [Green Version]
  94. Konsolakis, M.; Yentekakis, I.V.; Palermo, A.; Lambert, R. Optimal Promotion by Rubidium of the CO + NO Reaction Over Pt/γ-Al2O3 catalysts. Appl. Catal. B Environ. 2001, 33, 293–302. [Google Scholar] [CrossRef]
  95. Li, L.; Zhang, F.; Guan, N.; Schreier, E.; Richter, M. NO Selective Reduction by Hydrogen on Potassium Titanate Supported Palladium Catalyst. Catal. Commun. 2008, 9, 1827–1832. [Google Scholar] [CrossRef]
  96. Son, I.H.; Kim, M.C.; Koh, H.L.; Kim, K.-L. On the Promotion of Ag/γ-Al2O3 by Cs for the SCR of NO by C3H6. Catal. Lett. 2001, 75, 191–197. [Google Scholar] [CrossRef]
  97. Feng, B.; Lu, G.; Wang, Y.; Guo, Y.; Guo, Y. Promoting Role of Potassium on the Catalytic Performance of Copper Oxide for the Reduction of NO by Activated Carbon. Chin. J. Catal. 2011, 32, 853–861. [Google Scholar] [CrossRef]
  98. Wang, X.; Maeda, N.; Meier, D.M.; Baiker, A. Potassium Titanate Nanobelts: A Unique Support for Au and AuRh Nanoparticles in the Catalytic Reduction of NO with CO. ChemCatChem 2021, 13, 438–444. [Google Scholar] [CrossRef]
  99. Gholami, Z.; Luo, G.; Gholami, F.; Yang, F. Recent Advances in Selective Catalytic Reduction of Nox by Carbon Monoxide for Flue Gas Cleaning Process: A Review. Catal. Rev. 2021, 63, 68–119. [Google Scholar] [CrossRef]
  100. Konsolakis, M.; Yentekakis, I.V. NO Reduction by Propene or CO Over Alkali-promoted Pd/YSZ Catalysts. J. Hazard. Mater. 2007, 149, 619–624. [Google Scholar] [CrossRef] [PubMed]
  101. Rao, K.N.; Ha, H.P. SO2 Promoted Alkali Metal Doped Ag/Al2O3 Catalysts for CH4-SCR of NOx. Appl. Catal. A Gen. 2012, 433–434, 162–169. [Google Scholar] [CrossRef]
  102. Morales, C.; López, N.; Aguila, G.; Araya, P.; Scott, F.; Vergara-Fernández, A.; Guerrero, S. Simultaneous use of CO and Naphthalene for the Reduction of NO on Potassium Promoted Copper Catalyst Supported on Ce/TiO2-SiO2 and in the Presence of Oxygen. Mater. Chem. Phys. 2019, 222, 294–299. [Google Scholar] [CrossRef]
  103. Wang, Q.; Park, S.Y.; Choi, J.S.; Chung, J.S. Co/KxTi2O5 Catalysts Prepared by Ion Exchange Method for NO Oxidation to NO2. Appl. Catal. B Environ. 2008, 79, 101–107. [Google Scholar] [CrossRef]
  104. Yentekakis, I.; Lambert, R.; Tikhov, M.; Konsolakis, M.; Kiousis, V. Promotion by Sodium in Emission Control Catalysis: A Kinetic and Spectroscopic Study of the Pd-Catalyzed Reduction of NO by Propene. J. Catal. 1998, 176, 82–92. [Google Scholar] [CrossRef] [Green Version]
  105. MacLeod, N.; Isaac, J.; Lambert, R.M. Sodium Promotion of Pd/γ-Al2O3 Catalysts Operated under Simulated “Three-Way” Conditions. J. Catal. 2001, 198, 128–135. [Google Scholar] [CrossRef]
  106. Yentekakis, I.; Tellou, V.; Botzolaki, G.; Rapakousios, I. A Comparative Study of the C3H6+NO+O2, C3H6+O2 and NO+O2 Reactions in Excess Oxygen Over Na-Modified Pt/γ-Al2O3 Catalysts. Appl. Catal. B Environ. 2005, 56, 229–239. [Google Scholar] [CrossRef]
  107. Xie, Y.; Galvez, M.E.; Matynia, A.; Da Costa, P. Experimental Investigation on the Influence of the Presence of Alkali Compounds on the Performance of a Commercial Pt–Pd/Al2O3 Diesel Oxidation Catalyst. Clean Technol. Environ. Policy 2017, 20, 715–725. [Google Scholar] [CrossRef]
  108. Cao, Y.; Wang, F.; Wei, S.; Yang, W.; Zhou, Y. The Role of Potassium in the Activation of Oxygen to Promote Nitric Oxide Oxidation on Honeycomb-like h-BN(001) Surfaces. Phys. Chem. Chem. Phys. 2018, 20, 26777–26785. [Google Scholar] [CrossRef] [PubMed]
  109. Yang, Z.; Wang, J.; Yu, X. Density Functional Theory Studies on the Adsorption, Diffusion and Dissociation of O2 on Pt(111). Phys. Lett. A 2010, 374, 4713–4717. [Google Scholar] [CrossRef]
  110. Tang, X.; Gao, F.; Xiang, Y.; Yi, H.; Zhao, S.; Liu, X.; Li, Y. Effect of Potassium-Precursor Promoters on Catalytic Oxidation Activity of Mn-CoOx Catalysts for NO Removal. Ind. Eng. Chem. Res. 2015, 54, 9116–9123. [Google Scholar] [CrossRef]
  111. Shang, Z.; Sun, M.; Che, X.; Wang, W.; Wang, L.; Cao, X.; Zhan, W.; Guo, Y.; Guo, Y.; Lu, G. The Existing States of Potassium Species in K-Doped Co3O4 Catalysts and Their Influence on the Activities for NO and Soot Oxidation. Catal. Sci. Technol. 2017, 7, 4710–4719. [Google Scholar] [CrossRef]
  112. Adjimi, S.; García-Vargas, J.; Díaz, J.; Retailleau, L.; Gil, S.; Pera-Titus, M.; Guo, Y.; Giroir-Fendler, A. Highly Efficient and Stable Ru/K-OMS-2 Catalyst for NO Oxidation. Appl. Catal. B Environ. 2017, 219, 459–466. [Google Scholar] [CrossRef]
  113. Gálvez, M.E.; Ascaso, S.; Moliner, R.; Lázaro, M.J. Me (Cu, Co, V)-K/Al2O3 Supported Catalysts for the Simultaneous Removal of Soot and Nitrogen Oxides from Diesel Exhausts. Chem. Eng. Sci. 2013, 87, 75–90. [Google Scholar] [CrossRef]
  114. Aneggi, E.; De Leitenburg, C.; Dolcetti, G.; Trovarelli, A. Diesel Soot Combustion Activity of Ceria Promoted with Alkali Metals. Catal. Today 2008, 136, 3–10. [Google Scholar] [CrossRef]
  115. Zhang, Y.; Su, Q.; Li, Q.; Wang, Z.; Gao, X.; Zhang, Z. Determination of Mechanism for Soot Oxidation with NO on Potassium Supported Mg-Al Hydrotalcite Mixed Oxides. Chem. Eng. Technol. 2011, 34, 1864–1868. [Google Scholar] [CrossRef]
  116. Jimenez, R.; García, X.; Lopez, T.; Gordon, A. Catalytic Combustion of Soot. Effects of Added Alkali Metals on CaO–MgO Physical Mixtures. Fuel Process. Technol. 2008, 89, 1160–1168. [Google Scholar] [CrossRef]
  117. Weng, D.; Li, J.; Wu, X.; Lin, F. Promotional Effect of Potassium on Soot Oxidation Activity and SO2-Poisoning Resistance of Cu/CeO2 catalyst. Catal. Commun. 2008, 9, 1898–1901. [Google Scholar] [CrossRef]
  118. Gross, M.S.; Ulla, M.A.; Querini, C.A. Catalytic Oxidation of Diesel Soot: New Characterization and Kinetic Evidence Related to the Reaction Mechanism on K/CeO2 Catalyst. Appl. Catal. A Gen. 2009, 360, 81–88. [Google Scholar] [CrossRef]
  119. Hirabayashi, D.; Hashikawa, H.; Suzuki, K. Low Temperature Soot Oxidation via the Melting Process of Alkaline Salts. Waste Manag. Environ. IV 2008, 109, 847–856. [Google Scholar] [CrossRef] [Green Version]
  120. Yang, L.; Zhang, C.; Shu, X.; Yue, T.; Wang, S.; Deng, Z. The Mechanism of Pd, K Co-doping on Mg–Al Hydrotalcite for Simultaneous Removal of Diesel Soot and NOx in SO2-Containing Atmosphere. Fuel 2019, 240, 244–251. [Google Scholar] [CrossRef]
  121. Cao, C.; Xing, L.; Yang, Y.; Tian, Y.; Ding, T.; Zhang, J.; Hu, T.; Zheng, L.; Li, X. Diesel Soot Elimination over Potassium-Promoted Co3O4 Nanowires Monolithic Catalysts under Gravitation Contact Mode. Appl. Catal. B Environ. 2017, 218, 32–45. [Google Scholar] [CrossRef]
  122. Rinkenburger, A.; Toriyama, T.; Yasuda, K.; Niessner, R. Catalytic Effect of Potassium Compounds in Soot Oxidation. ChemCatChem 2017, 9, 3513–3525. [Google Scholar] [CrossRef]
  123. Jiménez, R.; García, X.; Gordon, A.L. CaO-MgO Catalysts for Soot Combustion: KnO3 as Source for Doping with Potassium. J. Chil. Chem. Soc. 2005, 50, 651–665. [Google Scholar] [CrossRef]
  124. Jiménez, R.; García, X.; Cellier, C.; Ruiz, P.; Gordon, A.L. Soot Combustion with K/MgO as Catalyst. Appl. Catal. A Gen. 2006, 297, 125–134. [Google Scholar] [CrossRef]
  125. Pecchi, G.; Cabrera, B.; Buljan, A.; Delgado, E.; Gordon, A.; Jimenez, R. Catalytic Oxidation of Soot over Alkaline Niobates. J. Alloy. Compd. 2013, 551, 255–261. [Google Scholar] [CrossRef]
  126. McKee, D.W. Mechanisms of the Alkali Metal Catalysed Gasification of Carbon. Fuel 1983, 62, 170–175. [Google Scholar] [CrossRef]
  127. Jiménez, R.; García, X.; Cellier, C.; Ruiz, P.; Gordon, A.L. Soot Combustion with K/MgO as Catalyst II. Effect of K-precursor. Appl. Catal. A Gen. 2006, 314, 81–88. [Google Scholar] [CrossRef]
  128. Li, Q.; Meng, M.; Zou, Z.-Q.; Li, X.-G.; Zha, Y.-Q. Simultaneous Soot Combustion and Nitrogen Oxides Storage on Potassium-Promoted Hydrotalcite-Based CoMgAlO Catalysts. J. Hazard. Mater. 2009, 161, 366–372. [Google Scholar] [CrossRef]
  129. Tikhomirov, K.; Kröcher, O.; Wokaun, A. Influence of Potassium Doping on the Activity and the Sulfur Poisoning Resistance of Soot Oxidation Catalysts. Catal. Lett. 2006, 109, 49–53. [Google Scholar] [CrossRef] [Green Version]
  130. Zhang, Y.; Zou, X. The Catalytic Activities and Thermal Stabilities of Li/Na/K Carbonates for Diesel Soot Oxidation. Catal. Commun. 2007, 8, 760–764. [Google Scholar] [CrossRef]
  131. Li, C.; Soh, K.C.K.; Wu, P. Formability of ABO3 Perovskites. J. Alloy. Compd. 2004, 372, 40–48. [Google Scholar] [CrossRef]
  132. Fino, D.; Russo, N.; Saracco, G.; Specchia, V. The Role of Suprafacial Oxygen in Some Perovskites for the Catalytic Combustion of Soot. J. Catal. 2003, 217, 367–375. [Google Scholar] [CrossRef]
  133. Fino, D.; Russo, N.; Badini, C.; Saracco, G.; Specchia, V. Effect of Active Species Mobility on Soot-Combustion over CS-V Catalysts. AIChE J. 2003, 49, 2173–2180. [Google Scholar] [CrossRef]
  134. Neri, G. K- and Cs-FeV/Al2O3 Soot Combustion Catalysts for Diesel Exhaust Treatment. Appl. Catal. B Environ. 2003, 42, 381–391. [Google Scholar] [CrossRef]
  135. Stanmore, B.; Brilhac, J.; Gilot, P. The Oxidation of Soot: A Review of Experiments, Mechanisms and Models. Carbon 2001, 39, 2247–2268. [Google Scholar] [CrossRef]
  136. Badini, C.; Saracco, G.; Russo, N.; Specchia, V. A Screening Study on the Activation Energy of Vanadate-Based Catalysts for Diesel Soot Combustion. Catal. Lett. 2000, 69, 207–215. [Google Scholar] [CrossRef]
  137. Weng, D.; Li, J.; Wu, X.; Si, Z. Modification of CeO2-ZrO2 Catalyst by Potassium for NOx-Assisted Soot Oxidation. J. Environ. Sci. 2011, 23, 145–150. [Google Scholar] [CrossRef]
  138. Zhang, Z.; Mou, Z.; Yu, P.; Zhang, Y.; Ni, X. Diesel Soot Combustion on Potassium Promoted Hydrotalcite-Based Mixed Oxide Catalysts. Catal. Commun. 2007, 8, 1621–1624. [Google Scholar] [CrossRef]
  139. Fan, Z.; Shi, J.; Zhang, Z.; Chen, M.; Shangguan, W. Promotion Effect of Potassium Carbonate on Catalytic Activity of Co3 O4 for Formaldehyde Removal. J. Chem. Technol. Biotechnol. 2018, 93, 3562–3568. [Google Scholar] [CrossRef]
  140. Liu, H.; Jia, A.; Yuwang; Luo, M.; Lu, J. Enhanced CO Oxidation over Potassium-Promoted Pt/Al2O3 Catalysts: Kinetic and Infrared Spectroscopic Study. Chin. J. Catal. 2015, 36, 1976–1986. [Google Scholar] [CrossRef]
  141. Tanaka, H.; Kuriyama, M.; Ishida, Y.; Ito, S.-I.; Tomishige, K.; Kunimori, K. Preferential CO Oxidation in Hydrogen-Rich Stream over Pt Catalysts Modified with Alkali Metals Part I. Catalytic performance. Appl. Catal. A Gen. 2008, 343, 117–124. [Google Scholar] [CrossRef] [Green Version]
  142. Kuriyama, M.; Tanaka, H.; Ito, S.-I.; Kubota, T.; Miyao, T.; Naito, S.; Tomishige, K.; Kunimori, K. Promoting Mechanism of Potassium in Preferential CO Oxidation on Pt/Al2O3. J. Catal. 2007, 252, 39–48. [Google Scholar] [CrossRef] [Green Version]
  143. Derrouiche, S.; Gravejat, P.; Bassou, B.; Bianchi, D. Impact of Potassium on the Heats of Adsorption of Adsorbed CO Species on Supported Pt Particles by Using the AEIR Method. Appl. Surf. Sci. 2007, 253, 5894–5898. [Google Scholar] [CrossRef]
  144. Niu, T.; Wang, C.; Zhang, L.; Liu, Y. Potassium Promoted Ru/Meso-Macroporous SiO2 Catalyst for the Preferential Oxidation of CO in H2-Rich Gases. Int. J. Hydrog. Energy 2013, 38, 7801–7810. [Google Scholar] [CrossRef]
  145. Pavlenko, N.; Kostrobij, P.; Suchorski, Y.; Imbihl, R. Alkali Metal Effect on Catalytic Co Oxidation on a Transition Metal Surface: A Lattice-Gas Model. Surf. Sci. 2001, 489, 29–36. [Google Scholar] [CrossRef]
  146. Yakovkin, I.; Chernyi, V.; Naumovets, A. Oxidation of CO on Li-precovered Pt. Surf. Sci. 1999, 442, 81–89. [Google Scholar] [CrossRef]
  147. Zhang, Y.; Wu, S.; Jin, L.; Xie, D. Development and Performance of a K–Pt/γ-Al 2 O 3 Catalyst for the Preferential Oxidation of CO in Hydrogen-Rich Synthesis Gas. Int. J. Hydrog. Energy 2014, 39, 18668–18674. [Google Scholar] [CrossRef]
  148. Liotta, L.F.; Martin, G.A.; Deganello, G. The Influence of Alkali Metal Ions in the Chemisorption of CO and CO2on Supported Palladium Catalysts: A Fourier Transform Infrared Spectroscopic Study. J. Catal. 1996, 164, 322–333. [Google Scholar] [CrossRef]
  149. Solymosi, F.; Pásztor, M.; Rákhely, G. Infrared Studies of the Effects of Promoters on CO-induced Structural Changes in Rh. J. Catal. 1988, 110, 413–415. [Google Scholar] [CrossRef]
  150. Dai, C.H.; Worley, S.D. Effects of Potassium on Carbon Monoxide Methanation over Supported Rhodium Films. J. Phys. Chem. 1986, 90, 4219–4221. [Google Scholar] [CrossRef]
  151. Rodriguez, J.A.; Grinter, D.C.; Ramírez, P.J.; Stacchiola, D.J.; Senanayake, S.D. High Activity of Au/K/TiO2(110) for CO Oxidation: Alkali-Metal-Enhanced Dispersion of Au and Bonding of CO. J. Phys. Chem. C 2018, 122, 4324–4330. [Google Scholar] [CrossRef]
  152. Erikat, I.A. First Principle Study on Alkali Metals Promotion of CO Oxidation over Ir(100). Phys. Status Solidi 2016, 253, 983–989. [Google Scholar] [CrossRef]
  153. Minemura, Y.; Ito, S.-I.; Miyao, T.; Naito, S.; Tomishige, K.; Kunimori, K. Preferential CO Oxidation Promoted by the Presence of H2 over K–Pt/Al2O3. Chem. Commun. 2005, 1429–1431. [Google Scholar] [CrossRef]
  154. Tanaka, H.; Ito, S.-I.; Kameoka, S.; Tomishige, K.; Kunimori, K. Promoting Effect of Potassium in Selective Oxidation of CO in Hydrogen-Rich Stream on Rh Catalysts. Catal. Commun. 2003, 4, 1–4. [Google Scholar] [CrossRef]
  155. Ang, M.L.; Oemar, U.; Saw, E.T.; Mo, L.; Kathiraser, Y.; Chia, B.H.; Kawi, S. Highly Active Ni/xNa/CeO2 Catalyst for the Water–Gas Shift Reaction: Effect of Sodium on Methane Suppression. ACS Catal. 2014, 4, 3237–3248. [Google Scholar] [CrossRef]
  156. Gluhoi, A.C.; Tang, X.; Marginean, P.; Nieuwenhuys, B.E. Characterization and Catalytic Activity of Unpromoted and Alkali (earth)-Promoted Au/Al2O3 Catalysts for Low-Temperature CO Oxidation. Top. Catal. 2006, 39, 101–110. [Google Scholar] [CrossRef]
  157. Tanaka, H.; Ito, S.-I.; Kameoka, S.; Tomishige, K.; Kunimori, K. Catalytic Performance of K-promoted Rh/USY Catalysts in Preferential Oxidation of CO in Rich Hydrogen. Appl. Catal. A Gen. 2003, 250, 255–263. [Google Scholar] [CrossRef]
  158. Sheffer, G. Potassium’s Promotional Effect of Unsupported Copper Catalysts for Methanol Synthesis. J. Catal. 1989, 115, 376–387. [Google Scholar] [CrossRef]
  159. Farkas, A.P.; Solymosi, F. Activation and Reactions of CO2 on a K-Promoted Au(111) Surface. J. Phys. Chem. C 2009, 113, 19930–19936. [Google Scholar] [CrossRef]
  160. Dongil, A.B.; Bachiller-Baeza, B.; Castillejos, E.; Escalona, N.; Guerrero-Ruiz, A.; Rodríguez-Ramos, I. The Promoter Effect of Potassium in CuO/CeO2 Systems Supported on Carbon Nanotubes and Graphene for the CO-PROX Reaction. Catal. Sci. Technol. 2016, 6, 6118–6127. [Google Scholar] [CrossRef]
  161. Kaplin, I.Y.; Lokteva, E.S.; Golubina, E.V.; Maslakov, K.I.; Chernyak, S.A.; Lunin, V.V. Promoting Effect of Potassium and Calcium Additives to Cerium–Zirconium Oxide Catalysts for the Complete Oxidation of Carbon Monoxide. Kinet. Catal. 2017, 58, 585–592. [Google Scholar] [CrossRef]
  162. Minemura, Y.; Kuriyama, M.; Ito, S.-I.; Tomishige, K.; Kunimori, K. Additive Effect of Alkali Metal Ions on Preferential CO Oxidation Over Pt/Al2O3. Catal. Commun. 2006, 7, 623–626. [Google Scholar] [CrossRef]
  163. Pedrero, C.; Waku, T.; Iglesia, E. Oxidation of CO in H2–CO Mixtures Catalyzed by Platinum: Alkali Effects on Rates and Selectivity. J. Catal. 2005, 233, 242–255. [Google Scholar] [CrossRef]
  164. Ito, S.-I.; Tanaka, H.; Minemura, Y.; Kameoka, S.; Tomishige, K.; Kunimori, K. Selective CO Oxidation in H2-Rich Gas over K2CO3-promoted Rh/SiO2 Catalysts: Effects of Preparation Method. Appl. Catal. A Gen. 2004, 273, 295–302. [Google Scholar] [CrossRef]
  165. Tanaka, H.; Kuriyama, M.; Ishida, Y.; Ito, S.-I.; Kubota, T.; Miyao, T.; Naito, S.; Tomishige, K.; Kunimori, K. Preferential CO Oxidation in Hydrogen-Rich Stream over PT Catalysts Modified with Alkali Metals Part II. Catalyst characterization and role of alkali metals. Appl. Catal. A Gen. 2008, 343, 125–133. [Google Scholar] [CrossRef] [Green Version]
  166. Bugyi, L.; Solymosi, F. Effects of Potassium on the Chemisorption of CO on the Mo2C/Mo(100) Surface. J. Phys. Chem. B 2001, 105, 4337–4342. [Google Scholar] [CrossRef]
  167. Liu, B.; Huang, T.; Zhang, Z.; Wang, Z.; Zhang, Y.; Li, J. The Effect of the Alkali Additive on the Highly Active Ru/C Catalyst for Water Gas Shift Reaction. Catal. Sci. Technol. 2014, 4, 1286–1292. [Google Scholar] [CrossRef]
  168. Nagai, M.; Zahidul, A.M.; Kunisaki, Y.; Aoki, Y. Water–Gas Shift Reactions on Potassium- and Zirconium-Promoted Cobalt Molybdenum Carbide Catalysts. Appl. Catal. A Gen. 2010, 383, 58–65. [Google Scholar] [CrossRef]
  169. Rodriguez, J.A.; Remensal, E.R.; Ramírez, P.J.; Orozco, I.; Liu, Z.; Graciani, J.; Senanayake, S.D.; Sanz, J.F. Water–Gas Shift Reaction on K/Cu(111) and Cu/K/TiO2(110) Surfaces: Alkali Promotion of Water Dissociation and Production of H2. ACS Catal. 2019, 9, 10751–10760. [Google Scholar] [CrossRef]
  170. Evin, H.N.; Jacobs, G.; Ruiz-Martinez, J.; Graham, U.M.; Dozier, A.; Thomas, G.; Davis, B.H. Low Temperature Water–Gas Shift/Methanol Steam Reforming: Alkali Doping to Facilitate the Scission of Formate and Methoxy C–H Bonds over Pt/ceria Catalyst. Catal. Lett. 2008, 122, 9–19. [Google Scholar] [CrossRef]
  171. Maneerung, T.; Hidajat, K.; Kawi, S. K-doped LaNiO3 Perovskite for High-Temperature Water-Gas Shift of Reformate Gas: Role of Potassium on Suppressing Methanation. Int. J. Hydrog. Energy 2017, 42, 9840–9857. [Google Scholar] [CrossRef]
  172. Kaftan, A.; Kusche, M.; Laurin, M.; Wasserscheid, P.; Libuda, J. KOH-promoted Pt/Al2O3 Catalysts for Water Gas Shift and Methanol Steam Reforming: An Operando DRIFTS-MS Study. Appl. Catal. B Environ. 2017, 201, 169–181. [Google Scholar] [CrossRef]
  173. Zhai, Y.; Pierre, D.; Si, R.; Deng, W.; Ferrin, P.; Nilekar, A.U.; Peng, G.; Herron, J.A.; Bell, D.C.; Saltsburg, H.; et al. Alkali-Stabilized Pt-OHx Species Catalyze Low-Temperature Water-Gas Shift Reactions. Science 2010, 329, 1633–1636. [Google Scholar] [CrossRef]
  174. Fu, Q.; Saltsburg, H.; Flytzani-Stephanopoulos, M. Active Nonmetallic Au and Pt Species on Ceria-Based Water-Gas Shift Catalysts. Science 2003, 301, 935–938. [Google Scholar] [CrossRef] [PubMed]
  175. Pierre, D.; Deng, W.; Flytzani-Stephanopoulos, M. The Importance of Strongly Bound Pt–CeOx Species for the Water-gas Shift Reaction: Catalyst Activity and Stability Evaluation. Top. Catal. 2007, 46, 363–373. [Google Scholar] [CrossRef]
  176. Pazmiño, J.H.; Shekhar, M.; Williams, W.D.; Akatay, M.C.; Miller, J.T.; Delgass, W.N.; Ribeiro, F.H. Metallic Pt as Active Sites for the Water–Gas Shift Reaction on Alkali-Promoted Supported Catalysts. J. Catal. 2012, 286, 279–286. [Google Scholar] [CrossRef]
  177. Kantschewa, M. Nature and Properties of a Potassium-Promoted NiMo/Al2O3 Water Gas Shift Catalyst. J. Catal. 1984, 87, 482–496. [Google Scholar] [CrossRef]
  178. Majima, T.; Kono, E.; Ogo, S.; Sekine, Y. Pre-Reduction and K Loading Effects on Noble Metal Free Co-system Catalyst for Water Gas Shift Reaction. Appl. Catal. A Gen. 2016, 523, 92–96. [Google Scholar] [CrossRef]
  179. Kordulis, C.; Voliotis, S.; Lycourghiotis, A. Molybdena Catalysts Prepared on Modified Carriers: Regulation of the Symmetry and Valence of the Molybdenum Species Formed on γ-Al2O3 Modified with Alkali Cations. J. Less Common Met. 1982, 84, 187–200. [Google Scholar] [CrossRef]
  180. Zhu, X.; Shen, M.; Lobban, L.L.; Mallinson, R.G. Structural Effects of Na Promotion for High Water Gas Shift Activity on Pt–Na/TiO2. J. Catal. 2011, 278, 123–132. [Google Scholar] [CrossRef]
  181. Ranjbar, A.; Irankhah, A.; Aghamiri, S.F. Catalytic Activity of Rare Earth and Alkali Metal Promoted (Ce, La, Mg, K) Ni/Al2O3 Nanocatalysts in Reverse Water Gas Shift Reaction. Res. Chem. Intermed. 2019, 45, 5125–5141. [Google Scholar] [CrossRef]
  182. Porosoff, M.D.; Baldwin, J.W.; Peng, X.; Mpourmpakis, G.; Willauer, H.D. Potassium-Promoted Molybdenum Carbide as a Highly Active and Selective Catalyst for CO2 Conversion to CO. ChemSusChem 2017, 10, 2408–2415. [Google Scholar] [CrossRef]
  183. Prasad, P.S.S.; Bae, J.W.; Jun, K.-W.; Lee, K.-W. Fischer–Tropsch Synthesis by Carbon Dioxide Hydrogenation on Fe-Based Catalysts. Catal. Surv. Asia 2008, 12, 170–183. [Google Scholar] [CrossRef]
  184. Kotarba, A.; Barański, A.; Hodorowicz, S.; Sokołowski, J.; Szytuła, A.; Holmlid, L. Stability and Excitation of Potassium Promoter in Iron Catalysts—the Role of KFeO2 and KAlO2 Phases. Catal. Lett. 2000, 67, 129–134. [Google Scholar] [CrossRef]
  185. Loiland, J.A.; Wulfers, M.J.; Marinkovic, N.S.; Lobo, R.F. Fe/γ-Al2O3 and Fe–K/γ-Al2O3 as Reverse Water-Gas Shift Catalysts. Catal. Sci. Technol. 2016, 6, 5267–5279. [Google Scholar] [CrossRef] [Green Version]
  186. Liang, B.; Duan, H.; Su, X.; Chen, X.; Huang, Y.; Chen, X.; Delgado, J.J.; Zhang, T. Promoting Role of Potassium in the Reverse Water Gas Shift Reaction on Pt/Mullite Catalyst. Catal. Today 2017, 281, 319–326. [Google Scholar] [CrossRef]
  187. Morse, J.R.; Juneau, M.; Baldwin, J.W.; Porosoff, M.D.; Willauer, H.D. Alkali Promoted Tungsten Carbide as a Selective Catalyst for the Reverse Water Gas Shift Reaction. J. CO2 Util. 2020, 35, 38–46. [Google Scholar] [CrossRef]
  188. Pistonesi, C.; Pronsato, M.E.; Bugyi, L.; Juan, A. The Adsorption of CO on Potassium Doped Molybdenum Carbide Surface: An Ab-Initio Study. Catal. Today 2012, 181, 102–107. [Google Scholar] [CrossRef]
  189. Chen, C.-S.; Cheng, W.-H.; Lin, S.-S. Study of Reverse Water Gas Shift Reaction by TPD, TPR and CO2 Hydrogenation over Potassium-promoted Cu/SiO2 Catalyst. Appl. Catal. A Gen. 2003, 238, 55–67. [Google Scholar] [CrossRef]
  190. Yang, X.; Su, X.; Chen, X.; Duan, H.; Liang, B.; Liu, Q.; Liu, X.; Ren, Y.; Huang, Y.; Zhang, T. Promotion Effects of Potassium on the Activity and Selectivity of Pt/Zeolite Catalysts for Reverse Water Gas Shift Reaction. Appl. Catal. B Environ. 2017, 216, 95–105. [Google Scholar] [CrossRef]
  191. Pastor-Pérez, L.; Shah, M.; Le Saché, E.; Reina, T.R. Improving Fe/Al2O3 Catalysts for the Reverse Water-Gas Shift Reaction: On the Effect of Cs as Activity/Selectivity Promoter. Catalysts 2018, 8, 608. [Google Scholar] [CrossRef] [Green Version]
  192. Wang, L.; Liu, H.; Chen, Y.; Zhang, R.; Yang, S. K-Promoted Co–CeO2 Catalyst for the Reverse Water–Gas Shift Reaction. Chem. Lett. 2013, 42, 682–683. [Google Scholar] [CrossRef]
  193. Williams, F.J.; Bird, D.P.C.; Palermo, A.; Santra, A.K.; Lambert, R.M. Mechanism, Selectivity Promotion, and New Ultraselective Pathways in Ag-Catalyzed Heterogeneous Epoxidation. J. Am. Chem. Soc. 2004, 126, 8509–8514. [Google Scholar] [CrossRef]
  194. Linic, S.; Barteau, M.A. On the Mechanism of Cs Promotion in Ethylene Epoxidation on Ag. J. Am. Chem. Soc. 2004, 126, 8086–8087. [Google Scholar] [CrossRef] [PubMed]
  195. Mross, W.D. Alkali Doping in Heterogeneous Catalysis. Catal. Rev. 1983, 25, 591–637. [Google Scholar] [CrossRef]
  196. Wang, Y.; Liu, H.-H.; Wang, S.-Y.; Luo, M.-F.; Lu, J.-Q. Remarkable Enhancement of Dichloromethane Oxidation over Potassium-Promoted Pt/Al2O3 Catalysts. J. Catal. 2014, 311, 314–324. [Google Scholar] [CrossRef]
  197. Zhang, C.; He, H.; Tanaka, K.-I. Catalytic Performance and Mechanism of a Pt/TiO2 Catalyst for the Oxidation of Formaldehyde at Room Temperature. Appl. Catal. B Environ. 2006, 65, 37–43. [Google Scholar] [CrossRef]
  198. Zhang, C.; Liu, F.; Zhai, Y.; Ariga, H.; Yi, N.; Liu, Y.; Asakura, K.; Flytzani-Stephanopoulos, M.; He, H. Alkali-Metal-Promoted Pt/TiO2 Opens a More Efficient Pathway to Formaldehyde Oxidation at Ambient Temperatures. Angew. Chem. Int. Ed. 2012, 51, 9628–9632. [Google Scholar] [CrossRef]
  199. Xu, J.; Ekblad, M.; Nishiyama, S.; Tsuruya, S.; Masai, M. Effect of Alkali Metals Added to Cu Ion-Exchanged Y-Type Zeolite Catalysts in the Gas-Phase Catalytic Oxidation of Benzyl Alcohol. J. Chem. Soc. Faraday Trans. 1998, 94, 473–479. [Google Scholar] [CrossRef]
  200. Kim, S.S.; Park, K.H.; Hong, S.C. A Study on HCHO Oxidation Characteristics at Room Temperature Using a Pt/TiO2 Catalyst. Appl. Catal. A Gen. 2011, 398, 96–103. [Google Scholar] [CrossRef]
  201. Luo, L.; Wang, S.; Fan, C.; Yang, L.; Wu, Z.; Qin, Z.; Zhu, H.; Fan, W.; Wang, J. Promoting Effect of Alkali Metal Cations on the Catalytic Performance of Pd/H-ZSM-5 in the Combustion of Lean Methane. Appl. Catal. A Gen. 2020, 602, 117678. [Google Scholar] [CrossRef]
  202. Wang, X.; Zhang, Q.; Yang, S.; Wang, Y. Iron-Catalyzed Propylene Epoxidation by Nitrous Oxide: Studies on the Effects of Alkali Metal Salts. J. Phys. Chem. B 2005, 109, 23500–23508. [Google Scholar] [CrossRef]
  203. Jirátová, K.; Mikulová, J.; Klempa, J.; Grygar, T.; Bastl, Z.; Kovanda, F. Modification of Co–Mn–Al Mixed Oxide with Potassium and Its Effect on Deep Oxidation of VOC. Appl. Catal. A Gen. 2009, 361, 106–116. [Google Scholar] [CrossRef]
  204. Sobczak, I.; Rydz, M.; Ziolek, M. The Effect of Alkali Metal on the Surface Properties of Potassium Doped Au-Beta Zeolites. Mater. Res. Bull. 2013, 48, 795–801. [Google Scholar] [CrossRef]
  205. Chen, M.; Zheng, X.-M. The Effect of K and Al over NiCo2O4 Catalyst on Its Character and Catalytic Oxidation of VOCs. J. Mol. Catal. A Chem. 2004, 221, 77–80. [Google Scholar] [CrossRef]
  206. Nakashima, D.; Ichihashi, Y.; Nishiyama, S.; Tsuruya, S. Promoted Partial Oxidation Activity of Alkali Metal Added-Co Catalysts Supported on NaY and NaUSY Zeolites in the Gas-Phase Catalytic Oxidation of Benzyl Alcohol. J. Mol. Catal. A Chem. 2006, 259, 108–115. [Google Scholar] [CrossRef]
  207. Li, Y.; Nakashima, D.; Ichihashi, Y.; Nishiyama, S.; Tsuruya, S. Promotion Effect of Alkali Metal Added to Impregnated Cobalt Catalysts in the Gas-Phase Catalytic Oxidation of Benzyl Alcohol. Ind. Eng. Chem. Res. 2004, 43, 6021–6026. [Google Scholar] [CrossRef]
  208. Avgouropoulos, G.; Oikonomopoulos, E.; Kanistras, D.; Ioannides, T. Complete Oxidation of Ethanol over Alkali-Promoted Pt/Al2O3 Catalysts. Appl. Catal. B Environ. 2006, 65, 62–69. [Google Scholar] [CrossRef]
  209. Filkin, N.C.; Tikhov, M.S.; Palermo, A.; Lambert, R.M. A Kinetic and Spectroscopic Study of thein SituElectrochemical Promotion by Sodium of the Platinum-Catalyzed Combustion of Propene. J. Phys. Chem. A 1999, 103, 2680–2687. [Google Scholar] [CrossRef]
  210. Vernoux, P. In-Situ Electrochemical Control of the Catalytic Activity of Platinum for the Propene Oxidation. Solid State Ionics 2004, 175, 609–613. [Google Scholar] [CrossRef]
  211. Petrolekas, P.D.; Brosda, S.; Vayenas, C.G. Electrochemical Promotion of Pt Catalyst Electrodes Deposited on Na3Zr2Si2 PO 12 during Ethylene Oxidation. J. Electrochem. Soc. 1998, 145, 1469–1477. [Google Scholar] [CrossRef]
  212. De Lucasconsuegra, A.; Dorado, F.; Valverde, J.L.; Karoum, R.; Vernoux, P. Low-Temperature Propene Combustion over Pt/K-β-Al2O3 Electrochemical Catalyst: Characterization, Catalytic Activity Measurements, and Investigation of the NEMCA Effect. J. Catal. 2007, 251, 474–484. [Google Scholar] [CrossRef]
  213. Aissat, A.; Siffert, S.; Courcot, D. Preparation of Alkali-M/ZrO2 (M = Co or Cu) for VOCs Oxidation in the Presence of NOx or carbonaceous particles. Stud. Surf. Sci. Catal. 2010, 175, 747–750. [Google Scholar] [CrossRef]
  214. Chmielarz, L.; Piwowarska, Z.; Rutkowska, M.; Wojciechowska, M.; Dudek, B.; Witkowski, S.; Michalik, M. Total Oxidation of Selected Mono-Carbon VOCs over Hydrotalcite Originated Metal Oxide Catalysts. Catal. Commun. 2012, 17, 118–125. [Google Scholar] [CrossRef]
  215. Aissat, A.; Siffert, S.; Courcot, D.; Cousin, R.; Aboukaïs, A. VOCs and Carbonaceous Particles Removal Assisted by NOx on Alkali0.15/ZrO2 and Csx–M0.1/ZrO2 Catalysts (M = Cu or Co). Comptes Rendus Chim. 2010, 13, 515–526. [Google Scholar] [CrossRef]
  216. Wu, Y.; Liu, M.; Ma, Z.; Xing, S.T. Effect of Alkali Metal Promoters on Natural Manganese Ore Catalysts for the Complete Catalytic Oxidation of O-Xylene. Catal. Today 2011, 175, 196–201. [Google Scholar] [CrossRef]
  217. Lee, J.H.; Koo, K.Y.; Jung, U.H.; Park, J.E.; Yoon, W.L. The Promotional Effect of K on the Catalytic Activity of Ni/MgAl2O4 for the Combined H2O and CO2 Reforming of Coke Oven Gas for Syngas Production. Korean J. Chem. Eng. 2016, 33, 3115–3120. [Google Scholar] [CrossRef]
  218. Barthos, R.; Széchenyi, A.; Koós, Á.; Solymosi, F. The Decomposition of Ethanol over Mo2C/Carbon Catalysts. Appl. Catal. A Gen. 2007, 327, 95–105. [Google Scholar] [CrossRef]
  219. Dömök, M.; Baán, K.; Kecskés, T.; Erdőhelyi, A. Promoting Mechanism of Potassium in the Reforming of Ethanol on Pt/Al2O3 Catalyst. Catal. Lett. 2008, 126, 49–57. [Google Scholar] [CrossRef]
  220. Ogo, S.; Shimizu, T.; Nakazawa, Y.; Mukawa, K.; Mukai, D.; Sekine, Y. Steam Reforming of Ethanol over K Promoted Co Catalyst. Appl. Catal. A Gen. 2015, 495, 30–38. [Google Scholar] [CrossRef]
  221. Pendem, C.; Sarkar, B.; Siddiqui, N.; Konathala, L.N.S.K.; Baskar, C.; Bal, R. K-Promoted Pt-Hydrotalcite Catalyst for Production of H2 by Aqueous Phase Reforming of Glycerol. ACS Sustain. Chem. Eng. 2017, 6, 2122–2131. [Google Scholar] [CrossRef]
  222. Luna, A.E.C.; Iriarte, M.E. Carbon Dioxide Reforming of Methane over a Metal Modified Ni-Al2O3 Catalyst. Appl. Catal. A Gen. 2008, 343, 10–15. [Google Scholar] [CrossRef]
  223. Juan-Juan, J.; Román-Martínez, M.; Illán-Gómez, M. Effect of Potassium Content in the Activity of K-Promoted Ni/Al2O3 Catalysts for the Dry Reforming of Methane. Appl. Catal. A Gen. 2006, 301, 9–15. [Google Scholar] [CrossRef]
  224. Hu, X.; Lu, G. Inhibition of Methane Formation in Steam Reforming Reactions through Modification of Ni Catalyst and the Reactants. Green Chem. 2009, 11, 724–732. [Google Scholar] [CrossRef]
  225. Darujati, A.R.; Thomson, W.J. Stability of Supported and Promoted-Molybdenum Carbide Catalysts in Dry-Methane Reforming. Appl. Catal. A Gen. 2005, 296, 139–147. [Google Scholar] [CrossRef]
  226. Osaki, T.; Mori, T. Role of Potassium in Carbon-Free CO2 Reforming of Methane on K-Promoted Ni/Al2O3 Catalysts. J. Catal. 2001, 204, 89–97. [Google Scholar] [CrossRef]
  227. Frusteri, F.; Arena, F.; Calogero, G.; Torre, T.; Parmaliana, A. Potassium-enhanced Stability of Ni/MgO Catalysts in the Dry-reforming of Methane. Catal. Commun. 2001, 2, 49–56. [Google Scholar] [CrossRef]
  228. Yu, X.; Zhang, S.; Wang, L.; Jiang, Q.; Li, S.; Tao, Z. Hydrogen Production from Steam Reforming of Kerosene Over Ni–La and Ni–La–K/Cordierite Catalysts. Fuel 2006, 85, 1708–1713. [Google Scholar] [CrossRef]
  229. Frusteri, F.; Freni, S.; Chiodo, V.; Spadaro, L.; Bonura, G.; Cavallaro, S. Potassium Improved Stability of Ni/MgO in the Steam Reforming of Ethanol for the Production of Hydrogen for MCFC. J. Power Sources 2004, 132, 139–144. [Google Scholar] [CrossRef]
  230. Iwasa, N.; Yamane, T.; Arai, M. Influence of Alkali Metal Modification and Reaction Conditions on the Catalytic Activity and Stability of Ni Containing Smec-Tite-Type Material for Steam Reforming of Acetic Acid. Int. J. Hydrog. Energy 2011, 36, 5904–5911. [Google Scholar] [CrossRef]
  231. Nandini, A.; Pant, K.; Dhingra, S. K-, CeO2-and Mn-promoted Ni/Al2O3 Catalysts for Stable CO2 Reforming of Methane. Appl. Catal. A Gen. 2005, 290, 166–174. [Google Scholar] [CrossRef]
  232. Frusteri, F.; Freni, S.; Chiodo, V.; Spadaro, L.; Di Blasi, O.; Bonura, G.; Cavallaro, S. Steam Reforming of Bio-Ethanol on Alkali-Doped Ni/MgO Catalysts: Hydrogen Production for MC Fuel Cell. Appl. Catal. A Gen. 2004, 270, 1–7. [Google Scholar] [CrossRef]
  233. Shi, H.; Ni, J.; Zheng, T.; Wang, X.; Wu, C.; Wang, Q. Remediation of Wastewater Contaminated by Antibiotics. A review. Environ. Chem. Lett. 2020, 18, 345–360. [Google Scholar] [CrossRef]
  234. Li, Y.; Guo, Y.; Zhu, T.; Ding, S. Adsorption and Desorption of SO2, NO and Chlorobenzene on Activated Carbon. J. Environ. Sci. 2016, 43, 128–135. [Google Scholar] [CrossRef]
  235. Guo, Y.; Li, Y.; Zhu, T.; Wang, J.; Ye, M. Modeling of Dioxin Adsorption on Activated Carbon. Chem. Eng. J. 2016, 283, 1210–1215. [Google Scholar] [CrossRef]
  236. Zhang, Y.; Li, C.; Zhu, Y.; Du, X.; Lyu, Y.; Li, S.; Zhai, Y. Insight into the Enhanced Performance of Toluene Removal from Simulated Flue Gas over Mn-Cu Oxides Modified Activated Coke. Fuel 2020, 276, 118099. [Google Scholar] [CrossRef]
Figure 1. Various pollutant emissions in China from 2006 to 2015 [8,9,10]. (a) the annual emissions of SO2, NOx and soot in china, and (b) the annual emissions of CO2, CO and NMVOCs in china.
Figure 1. Various pollutant emissions in China from 2006 to 2015 [8,9,10]. (a) the annual emissions of SO2, NOx and soot in china, and (b) the annual emissions of CO2, CO and NMVOCs in china.
Catalysts 11 00419 g001
Figure 2. Single-site mechanism of N2O decomposition [63]. Copyright 2017, Elsevier.
Figure 2. Single-site mechanism of N2O decomposition [63]. Copyright 2017, Elsevier.
Catalysts 11 00419 g002
Figure 3. Double-site mechanism of N2O decomposition [34]. Copyright 2008, Elsevier.
Figure 3. Double-site mechanism of N2O decomposition [34]. Copyright 2008, Elsevier.
Catalysts 11 00419 g003
Figure 4. Different alkali metals for N2O decomposition on cobalt-cerium composite oxide catalysts [41]. Copyright 2009, American Chemical Society.
Figure 4. Different alkali metals for N2O decomposition on cobalt-cerium composite oxide catalysts [41]. Copyright 2009, American Chemical Society.
Catalysts 11 00419 g004
Figure 5. The correlation between precursor and T50, reaction rate, and work function [35]. Copyright 2013, Elsevier.
Figure 5. The correlation between precursor and T50, reaction rate, and work function [35]. Copyright 2013, Elsevier.
Catalysts 11 00419 g005
Figure 6. Mechanism of K promoting catalytic dissociation of NO over Co-Mn-Al catalyst [73]. Copyright 2019, MDPI.
Figure 6. Mechanism of K promoting catalytic dissociation of NO over Co-Mn-Al catalyst [73]. Copyright 2019, MDPI.
Catalysts 11 00419 g006
Figure 7. K cladding model [81]. (a) an ideal compact core and shell, (b) a more realistic core and cracked shell, and (c) deactivated catalyst grain morphology. Copyright 2002, Elsevier.
Figure 7. K cladding model [81]. (a) an ideal compact core and shell, (b) a more realistic core and cracked shell, and (c) deactivated catalyst grain morphology. Copyright 2002, Elsevier.
Catalysts 11 00419 g007
Figure 8. The mechanism of soot oxidation on molten salt [119]. Copyright 2008, WIT Press.
Figure 8. The mechanism of soot oxidation on molten salt [119]. Copyright 2008, WIT Press.
Catalysts 11 00419 g008
Figure 9. The oxygen overflow mechanism of soot oxidation [120]. Copyright 2019, Elsevier.
Figure 9. The oxygen overflow mechanism of soot oxidation [120]. Copyright 2019, Elsevier.
Catalysts 11 00419 g009
Figure 10. A nitrite-ketene mechanism for soot oxidation by NO [115]. M stands for metal and “*” stands for the free carbon sites on soot. Copyright 2011, Wiley Online Library.
Figure 10. A nitrite-ketene mechanism for soot oxidation by NO [115]. M stands for metal and “*” stands for the free carbon sites on soot. Copyright 2011, Wiley Online Library.
Catalysts 11 00419 g010
Figure 11. Reaction mechanism for soot oxidation on KCo-NW catalysts under NO/O2 atmosphere [121]. Copyright 2017, Elsevier.
Figure 11. Reaction mechanism for soot oxidation on KCo-NW catalysts under NO/O2 atmosphere [121]. Copyright 2017, Elsevier.
Catalysts 11 00419 g011
Figure 12. Reaction network for soot oxidation on K/CeO2 catalysts [118]. Copyright 2009, Elsevier.
Figure 12. Reaction network for soot oxidation on K/CeO2 catalysts [118]. Copyright 2009, Elsevier.
Catalysts 11 00419 g012
Figure 13. Mechanism of improvement of CO2 selectivity over Ni-based catalysts [155]. Copyright 2014, ACS Publications.
Figure 13. Mechanism of improvement of CO2 selectivity over Ni-based catalysts [155]. Copyright 2014, ACS Publications.
Catalysts 11 00419 g013
Figure 14. The mechanism of alkali metals improving the selectivity of water-gas shift (WGS) [171]. (a) the reduced LaNiO3 perovskite catalyst and (b) the reduced K-doped LaNiO3 perovskite catalyst. Copyright 2017, Elsevier.
Figure 14. The mechanism of alkali metals improving the selectivity of water-gas shift (WGS) [171]. (a) the reduced LaNiO3 perovskite catalyst and (b) the reduced K-doped LaNiO3 perovskite catalyst. Copyright 2017, Elsevier.
Catalysts 11 00419 g014
Figure 15. Effect of alkali metals on selectivity of propylene oxidation over CuOx/SiO2‘ [16]. Copyright 2018, Elsevier.
Figure 15. Effect of alkali metals on selectivity of propylene oxidation over CuOx/SiO2‘ [16]. Copyright 2018, Elsevier.
Catalysts 11 00419 g015
Figure 16. Reaction pathway of H2 formation in glycerol reforming promoted by K [221]. Copyright 2017, American Chemical Society.
Figure 16. Reaction pathway of H2 formation in glycerol reforming promoted by K [221]. Copyright 2017, American Chemical Society.
Catalysts 11 00419 g016
Figure 17. The correlation of the promotion effect of K on the commonly used active centers and types of promotion effect for NOx, COx, and VOCs.
Figure 17. The correlation of the promotion effect of K on the commonly used active centers and types of promotion effect for NOx, COx, and VOCs.
Catalysts 11 00419 g017
Table 1. The content of various pollutants in various gases.
Table 1. The content of various pollutants in various gases.
ComponentsSintering Flue GasPelletizing Flue GasCoke Oven Flue GasMotor Vehicle Exhaust
NOx (mg/m3)200–400200–300300–1000500–4000
Soot (g/m3)1–51–100.02–0.12–6 *
CO (%)0.3–10.6–10.04–0.81–8
CO2 (%)3–71–1.54–56–16
NMVOCs (g/m3)0.5–56–1815–201–50
* This represents the content of diesel engines.
Table 12. Summary of the factors affecting the promotion effect of alkali metals.
Table 12. Summary of the factors affecting the promotion effect of alkali metals.
TypeContentPreparationPrecursorCalcination TemperatureCalcination timeCarrierReaction TemperatureAtmosphere
N2O decompositionImpregnationK2CO3-----
NO decomposition-co-precipitation----
Soot oxidationK------
CO oxidationKSequential impregnation---electron transfer effect--
WGSK-K2CO3---Not too high-
RWGS--------
VOCs oxidaitonK-KCl----stoichiometric ratio
Reforming-- ------
This symbol represents a positive correlation between activity and factor. This symbol represents a negative correlation between activity and factor. ▲ This symbol represents a volcanic type relationship between activity and factor.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xu, Z.; Li, Y.; Shi, H.; Lin, Y.; Wang, Y.; Wang, Q.; Zhu, T. Application Prospect of K Used for Catalytic Removal of NOx, COx, and VOCs from Industrial Flue Gas: A Review. Catalysts 2021, 11, 419. https://doi.org/10.3390/catal11040419

AMA Style

Xu Z, Li Y, Shi H, Lin Y, Wang Y, Wang Q, Zhu T. Application Prospect of K Used for Catalytic Removal of NOx, COx, and VOCs from Industrial Flue Gas: A Review. Catalysts. 2021; 11(4):419. https://doi.org/10.3390/catal11040419

Chicago/Turabian Style

Xu, Zhicheng, Yuran Li, Huimin Shi, Yuting Lin, Yan Wang, Qiang Wang, and Tingyu Zhu. 2021. "Application Prospect of K Used for Catalytic Removal of NOx, COx, and VOCs from Industrial Flue Gas: A Review" Catalysts 11, no. 4: 419. https://doi.org/10.3390/catal11040419

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop