Next Article in Journal
Recent Approaches to Chiral 1,4-Dihydropyridines and their Fused Analogues
Next Article in Special Issue
Application of Activated Carbon to Obtain Biodiesel from Vegetable Oils
Previous Article in Journal
Green Catalysts: Applied and Synthetic Photosynthesis
Previous Article in Special Issue
Enabling Technologies and Sustainable Catalysis in Biodiesel Preparation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Monolith Metal-Oxide-Supported Catalysts: Sorbent for Environmental Application

by
Kiman Silas
1,2,
Wan Azlina Wan Ab Karim Ghani
1,2,3,*,
Thomas Shean Yaw Choong
1 and
Umer Rashid
4,*
1
Department of Chemical and Environmental Engineering, Faculty of Engineering, Universiti Putra Malaysia, UPM Serdang, Selangor 43400, Malaysia
2
Sustainable Process Engineering Research Center (SPERC), Faculty of Engineering, Universiti Putra Malaysia, UPM Serdang, Selangor 43400, Malaysia
3
Institute of Plantations Studies, Universiti Putra Malaysia, UPM Serdang, Selangor 43400, Malaysia
4
Institute of Advanced Technology, Universiti Putra Malaysia, UPM Serdang, Selangor 43400, Malaysia
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(9), 1018; https://doi.org/10.3390/catal10091018
Submission received: 22 July 2020 / Revised: 7 August 2020 / Accepted: 8 August 2020 / Published: 4 September 2020
(This article belongs to the Special Issue Sustainable Catalysts for Biofuel Production)

Abstract

:
The emission of untreated environmental harmful gases such as sulfur and nitrogen oxide (SOx and NOx) emissions is considered old fashioned, since industries are compelled by governments and legislations to meet the minimum threshold before emitting such substances into the atmosphere. Numerous research has been done and is ongoing to come up with both cost-effective equipment and regenerable catalysts that are adsorbent—or with enhanced sorption capacity—and with safer disposal methods. This work presents the general idea of a monolith/catalyst for environmental application and the technicality for improving the surface area for fast and efficient adsorption–desorption reactions. The chemical reactions, adsorption kinetics, and other properties, including deactivation, regeneration, and the disposal of a catalyst in view of environmental application, are extensively discussed.

1. Introduction

Atmospheric pollutants include acid gases such as carbon dioxide (CO2), sulfur oxides (SOx), nitrogen oxides (NOx), volatile organic compounds (VOCs), heavy metals, fly ash, and particulate matter [1,2,3]. The emitted pollutants are as a result of industrial activities from coating industries, fluid catalytic cracking units (FCCU), circulating fluidized bed combustors (CFBC), combustion behaviors of incinerators, coal fuel, automotives, plastic, and power plants which give rise to deleterious health and environmental effects [4,5,6,7,8,9]. For example, the world consumption of about 5–6 billion metric tons of coal is reported annually, and by 2040, it is expected to reach 12,500 Mt [10]. Coal is the dominant fossil fuel currently in use that accounted for 38% of electricity generated in 2000, 17.5% hydropower, 17.3% natural gas, 16.8% nuclear, 9% oil, and 1.6% non-hydropower, yet coal is expected to remain the dominant fuel for power generation by 2020 [11].
Adsorption is a simple and cheap method for extracting process gases and vapors from the atmosphere, and the adsorption of contaminants by regenerable catalysts is becoming common in reducing the menace of air pollution [12]. Recently, the multipollutant simultaneous removal techniques used to include carbon-based material adsorption and nonthermal plasma (NTP) technologies [13]. In addition, attention has been paid to the chemical reaction of acid gases with a metal oxide (chemical adsorbent) or the actual sorption of solids (atomic adsorbent) [14]. The solid adsorbents commonly used in controlling flue gas pollutants are the non-regenerative (CaO and MgO) and regenerative (zeolites, silica gel, charcoal, etc.) adsorbents [15]. The final disposal of the exhausted catalyst is a paramount topic of research. Commonly, a layer of sealant, such as bitumen, polyethylene or concrete, is used to encapsulate catalyst residues that may cause environmental problems [16]. In terms of economy, the regeneration of the exhausted catalyst is preferred over the adsorbent replacement [17].
Numerous research has been done and is ongoing to come up with cost-effective equipment and regenerable catalysts that are adsorbent—or with enhanced sorption capacity—and with safer disposal methods. For example, some literatures [18,19,20,21,22,23,24], worked on the development of catalysts but have not considered the regeneration aspects. Other work has reported the regenerations of catalysts [3,25,26,27,28] but failed to give good performance after few regeneration cycles due to poor understanding of the catalyst development. Besides, most of the studies does not give attention to the acid gases as the adsorbates, meanwhile activated carbon (AC) is mostly used as the catalyst support [29,30,31,32]. However, Kiman et al. [33] recommended the support of metal oxides catalysts with monolith due to issues with AC, such as low adsorptivity, plugging, and difficulty in mass transport processes, which can be overcome with the monolith adsorbent. Therefore, there is a growing need for the review of studies of catalysts with support and regeneration for environmental protection and cost effectiveness.
This study reviewed the potential of monolith as a metal oxide catalyst support. Furthermore, the industrial application specifically in the SOx/NOx abatement from flue gas and the background of laboratory adsorptive equipment, basic chemistry of catalysis, application, and regeneration are discussed.

2. Monolith Structure

Monolith could be a metal, plastic, or ceramic structure with continuous unitary open structures without bends that hardly obstructs flow; it has parallel channels that extrude in various geometry (circular, square, rectangle, hexagonal etc.) [34,35,36,37]. A vehicle monolithic exhaust converter is reported to have the diameter of D + 0.1 m, while the monolith sections for treating flue gas of power plants has a large-diameter of D’2 m [36]. The structure is defined by width, wall thickness, void fraction (varies between 0.5 and 0.9 and is commonly expressed as the open front region, OFA), channel openings (dh), and cell density (ranges between 100 and 1200 channels per square inch cpsi) [38,39]. Silas et al. [3] gave the specification of a bare monolith (cordierite) they used in their study of flue gas cleaning as follow in Table 1:
Cordierite can be used as a catalyst support; however, it has a low specific surface area and weak catalyst-support interaction [40]. The specific surface area of cordierite is less than 1 m2 g−1, meanwhile coating can increase the specific surface area (to about 15–30 m2 g−1) and improve the dispersion of active components. Moreover, acidic treatment increases the surface area up to 200 m2 g−1 [35,40,41]. Washcoat (provides high surface area), substrate (gives the shape and mechanical properties), and active phase (the catalytic element) are considered as the three components of modified structured catalysts [42,43]. Lei et al. [41] found that metal oxide may be used as the active phase, γ-Al2O3 as the washcoat and cordierite as support in the application where dust and carbon are avoided while handling a large volume of flue gases. The impregnation and calcinations of catalyst/support are significant processes in the catalyst development.
The term calcination is referred to the heat-treatment beyond drying and without the formation of a liquid phase, it can be carried out in different atmospheres (N2, vacuum, etc.) and at temperatures higher than those used in the catalytic reaction or catalyst regeneration [43]. The process converts the impregnated metal precursor (i.e., Cu(NO3)2·3H2O to CuO) to the formation of finished metal oxide catalysts [14]. At a high calcination temperature, the activity of the catalyst will be low because some redundant oxides occupy the active sites and its dispersion on the support [44]. Figure 1 is the structure of the monolith.
Recent studies have focused on the use of monolithic adsorbents as catalyst support. Table 2 depicts the recent monoliths used in the literature.

3. Catalysts

Industrial catalysts are either heterogeneous (catalysts and reactants in distinct physical phases) or homogeneous (catalysts and reactants in the same phase) [49]. A heterogeneous catalyst is a material that has a relative amount of different components characterized by active species, surface area, physical and/or chemical promoters, size, supports, shape, and pore volume distribution, which give rise to an optimum catalyst with good selectivity, activity, lifetime, ease of regeneration, and low cost [43]. According to Teoh et al. [50], if the adsorption process occurs on a carbon-coated monolith’s surface, it is heterogeneous. Most of the industries prefer heterogeneous catalysts because of the advantages they offer, such as potential reusability, economic and environmental friendliness, stability, low toxicity, and ease of recovery from the reaction system [51,52,53,54,55,56,57,58].
Approximately 80% of the chemical products generated are manufactured through heterogeneous catalytic processes and generate global annual sales of around USD 1500 billion [51,52,53,54,55,56,57,58,59]. Furthermore, catalysis is used in the production of polymers, pharmaceuticals, cosmetics, foods, in new processes for the generation of clean energy and abating environmental pollutants [51]. Meanwhile, the homogeneous catalysts exhibit difficulty to recycle, complicated processes of post treatment, and enormous amounts of waste-water generation [53,54,55,60]. Supported metal oxides are known for unstable activity in the industrial applications due to H2O, hydrocarbons, O2, temperature, and acid gases deactivation; therefore, the development of catalytic materials that can be used for eliminating atmospheric pollutants is required [61,62].

3.1. Metal-Oxide-Supported Catalysts

Catalysts without support are easier to be deactivated than supported catalysts; therefore, the synthesis and supporting of nanophase metal oxides with small particle size and high surface area may provide the improvements needed for industrial catalysts [63,64], since low metal oxide loading on the support can tremendously enhance the catalyst’s performance [65]. Already, the technology of impregnating metal oxides (Mn2, O3, NiO, Co3O4, V2O5, CuO, etc.) and noble metals (Pt, Pd, and Rh) on structural adsorbents for application in stationary and automotive pollutants removal is established [14]. It is proven that the catalyst coating on monoliths enhances the structural properties like resistance to thermal shock and convenient separation from media [66].
Figure 2 shows the scanning electron microscopy and energy dispersed X-ray (SEM/EDX) spectra of monolith Co3O4 supported adsorbents. The energy dispersed X-ray (EDX) spectra showed the presence of Co3O4 of about 1.1% in the activated carbon monolith, while the presence of carbon (41.44%) is seen from the carbonization process development. The SEM image disclosed abundant pores with dispersed Co3O4 catalysts.
The catalytic activity and selectivity of a metal oxide is correlated with certain characteristics, such as oxygen non-stoichiometry, composition, surface area, volume, shape, reducibility, and pore structure [4,68]. In addition, the key considerations in preparation include the choice of active chemical composition, promotion of catalysts, active phase deposition methods, and oxidative and reductive treatments [4]. Furthermore, the support should have the following properties: high surface area and acidity, pore volume, thermal conductivity, and reactivity [69]. Figure 3 shows the procedure for monolith/Co3O4-based catalysts by impregnation of pore volume, precipitation of deposition, and methods of hydrothermal synthesis [70].
A good approach of developing the adsorbent surface is to manufacture a metal oxide material as an ordered porous structure [68]. An example is a monolithic cordierite-based catalyst (CuO/Al2O3) developed to remove SO2 and NO from flue gas at 350–400 °C simultaneously, but high reaction temperature and energy consumption were reported [44]. Boyano et al. [71] found that for adsorbent/catalyst to be economically competitive then it must perform at a low-temperature range of 100–300 °C.

3.2. Catalyst Loading

The good understanding of catalyst loading on the support is imperative because higher loading can result in active site blockage, metal sintering, and/or reduction in catalytic activity, and also, lower loading may lead to fewer active sites with low activity ascribed to incomplete coverage of the internal surface area [72,73,74]. Table 3 shows former catalyst loading on supports.
Figure 4 shows the alumina deposition using urea method, which illustrates the deposition precipitation technique. The pores in the monolith support were attached with the mixed metal oxides by the deposition precipitation technique, which acts as catalyst. Furthermore, the calcinations process ensured the consolidation of the support and the catalyst resulting into a unique composite for convenient application.

3.3. Catalyst Deactivation

Catalyst loss of activity with time is inevitable, and deactivation may occur in the case of cracking catalysts in seconds but may last for 5–10 years in iron catalyst synthesis or several years in human metabolism [16,77,78]. Common reasons for deactivation of the catalyst include: evaporation, carbonaceous deposit formation (coking), poisoning (phosphorous, presence of the SO2 in the flue gas), washout, reduction, accidental temperature rise, corrosion plugging of pores, and metal crystallite encapsulation; whereas, deactivation are classified as: poisoning, coking or fouling, sintering, and phase transformation [51,69,77,79]. Moulijn et al. [80], addresses how deactivation affects the rate of reaction since catalytic activity is proportional to the number of active sites as shown in the following equation:
K O = N T K i η
where, Ko and K i are the observed and intrinsic rate constants for the reactions per active site, NT is the total number of active sites, and η is the effectiveness factor.
Catalyst deactivation can be caused by the decreased in the number of active sites (NT decreases), decrease in the quality of the active sites ( K i decreases), and degradation in accessibility of the pore space (η decreases). Poisoning can cause a decrease in the number of active sites according to the following relation:
N T = N T ( 1 α )
where, α is the fraction of the sites poisoned. The relative loss of activity is given as:
Δ x = ( xo xt ) xo 100 %
where, Δ x is the relative loss of activity, x the conversion, and indices 0 and t stand for times of exposure [81]. The blockage of active sites as a result of poisoning are discussed in detail elsewhere [62,79,80,81,82,83]. Table 4 shows the types, treatment, and causes of catalyst deactivation.

3.4. Surface Area and Mechanism

The surface area value is a variable factor depending on how the isothermic adsorption is calculated and perceived. The model process of adsorption mechanisms is formulated for suitable equations, which predicts required parameters, such as surface area whose predicted values are used to validate adsorption equations and general behavior [84]. The porosity of materials is classed based on the basis of vapor adsorption behavior of each pore: [85,86,87,88,89]
  • micropores (d < 2 nm).
  • mesopores (2 nm < d < 50 nm).
  • macropores (50 nm < d < 200 nm).
Figure 5 shows the positioning of micro, meso, and macropores on a catalyst.
Monoliths are mainly mesoporous [3,85,90,91,92]. Previously, Hu et al. [93], found that mesopore volumes and surface area lie between the limits 0.1–0.5 cm/g and the range of 20–100 m/g and serves as the main transport arteries for the adsorbate. Because of their large specific surface area, highly ordered mesoporous structure and interconnected channels, mesoporous materials attract more attention [94]. It should be noticed that meso and macropores act as a passage to the micropores for adsorbing [95].
The high surface area and pore volume in a material are imparted by mesopores and micropores through numerous active sites and size selectivity for molecules, while macropores improve mass transfer in order to overcome diffusion difficulties found in the mesopores and micropores pores [68]. Furthermore, over the relative pressure range of about 0.05–0.35, the BET and Langmuir equations describe the adsorption process, whereas the Dubinin–Astakhov (DA) and Dubinin–Radushkevich (DR) equations interpret adsorption phenomena at much lower relative pressures [84].
Activation is a process of improving the adsorption capabilities of an adsorbent by various processes resulting in pore structure enhancement and a decrease in the weight of the substrate, referred to as burn-off [96]. The work of Lázaro et al. [91] specified that the use of CO2 as an activating agent can increase microporosity, mesoporosity, and macroporosity. The purpose of activation is to enlarge the diameters of the fine pores and create new pores [85]. Moreno-Castilla and Pérez-Cadenas [37] suggested activation with CO2 at 700–900 °C.
One of the most popular methods for studying the mechanism of a reaction is to investigate the influence of electronic factors on the rate of the reaction. Figure 6 shows the CO oxidation reaction on a CuMnOx catalyst surface.
The mechanism involves O2 adsorption to form O2* precursors, which split on a vicinal vacancy, while in the second mechanism, the O2 activation occurs via the kinetically applicable CO*-assisted O2 dissociation step without the specific concern of stable O2* precursors [97].

4. Adsorption Kinetics

Catalysis is based on changes in the kinetics of chemical reactions, while kinetics defines the relative rates of numerous competitive pathways available to reactants, while thermodynamics acts as a gateway to the most stable products [51]. Useful information on the behavior of adsorption could be obtained from thermodynamics parameters, such as entropy (DS), the free energy (DG), and enthalpy change (DH) [98]. Desorption is the reverse adsorption process by which adsorbed molecules removed from the adsorbent with about equal energy released during adsorption are supplied for desorption; such an energy requirement dominates the operating costs for some separations processes [99].
Adsorption refers to the change in the concentration of a substance at the interface compared to the adjacent phases, and the time dependence of adsorption on sorption surfaces is called the kinetics of adsorption [84,100]. The enrichment of ions, atoms, and molecules on the surface area of an interface occurs during the adsorption process. Adsorption kinetic studies provide useful information about the mechanism involved in the adsorbent–adsorbent reaction and the time required to complete the adsorption process, which is one of the requirements for choosing solid adsorbents for a particular application [101].
According to Rezaei and Webley [102], the volumetric working capacity, pressure drop, mass transfer, and thermal management governs the adsorbent performance of adsorptive gas separation processes for an adsorption isotherm. The adsorption kinetics is determined by the following steps: [100]
  • External diffusion; molecules diffuses towards the interface space from the bulk phase.
  • Internal diffusion; molecules diffuses inside the pores.
  • Surface diffusion; molecules diffuses in the surface phase.
  • Adsorption-desorption elementary processes.
Catalyst absorptivity is described by adsorption isotherms, which are characterized by constants that express the surface properties with models such as the Lagergren, Langmuir, and Freundlich used to fit the isotherms [98,103]. The important sorbent characteristics, including pore volume, pore size, specific surface area, and energy distribution, are described by the adsorption isotherm equations that deal with the physical adsorption of gases and vapor, whereas the interpretation of specific curves of the adsorption mechanism describes the efficiency of the sorbent in utilitarian processes, purification and separation [100]. The adsorption capacity of the catalyst on the adsorbent surface at the equilibrium state is given as: [50,104]
q e = ( C o C e ) V W
where qe is the adsorbed amount (mg/g), C o is the initial concentration (mg/L) and Ce is the equilibrium concentration (mg/L). V is the volume of the solution (L) and W is the mass of adsorbent used (mg), respectively. The adsorption capacity at predetermined time intervals is calculated by using following equation: [50,105]
q e = ( C o C t ) V w
where C t is the adsorbate concentration (mg/L) at time t(s). According to Chaudhary and Balomajumder [106], the removal efficiency (RE) can be calculated from:
RE = ( C o C t ) 100 % C o
Adsorption occurs in gas–solid systems at the outside and vicinity of solid surface structure while in absorption, so the molecules penetrate the surface layer and enter the bulk solid [107]. To calculate the adsorption capacity of a catalyst, Equation (6) is uniquely used in a liquid–solid interface while for a gas–solid interface the following equation is used:
q = y · F M 0 t b ( 1 C t C o )   d t
where Ct is the effluent concentration, Co is the influent concentration, and y, M, F, and tb are the mole fraction in the feed, amount of the catalyst (g), volumetric flow rate (mL/min), and breakthrough time (min), respectively. Several studies demonstrated the application of Equation (7) in calculating adsorption capacity. [3,92,98,108,109]. Adsorption capacity can have units of mg g−l or mmol g−l or cm3 g−1.

4.1. The Freundlich Isotherm Model

The Freundlich isotherm model was the earliest known empirical equation used for nonideal sorption that involves heterogeneous adsorption through a multilayer adsorption mechanism. The following postulations were considered in formulating the Freundlich model: [104]
(a) the adsorption layer is energetically heterogeneous and (b) the density of the adsorbed quantity increases. Freundlich model is defined as: [98]
q e = K f C e 1 / n
where, k f and n are Freundlich constants, 1/n is the heterogeneity factor and an indicator of adsorption capacity, becoming more heterogeneous as its value gets farther to one. The linearized form of Equation (9):
logq e = logk f + 1 / nC e
where kf and n could be determined from the slope and intercept of the linearized plot [50]. The correlation coefficient (r2) can be evaluated by plotting ln(qe) versus ln(Ce) [104].

4.2. The Langmuir Model

The Langmuir adsorption isotherm model was originally used to explain the chemistry of a pure gas on a nonporous surface and the physisorption of a pure gas or a gas mixture on micro and mesoporous adsorbents [110]. Langmuir experimental isotherm suggested an adsorption theory on a flat surface based on kinetic considerations, where there was a continuous process of molecular bombardment on the surface and subsequent desorption to maintain zero accumulation frequency on the surface in equilibrium [50]. It is also a model for the adsorption of gases onto solids using the following assumptions: [50,98,104]
(a) Only a monolayer of adsorbed material is formed at peak adsorption. Adsorbed molecules do not deposit on each other and the adsorption sites were the same. (b) The adsorbed molecules did not interact at constant temperature (c) The adsorbent layer was consistent with the same adsorption sites. (d) All adsorption by the same process occurred. The Langmuir is formulated as follows:
q e = ( q m KL C e ) ( 1 + KL C e )
where, q e is the equilibrium adsorption capacity (mg/g),   q m and KL are the maximum adsorption capacity to form a complete monolayer on the surface (mg/g), and the Langmuir constant related to the energy of adsorption (bonding energy of sorption in L/g), respectively. The linearized form of Langmuir equation is:
C e q e = 1 q m KL + C e q m
where q m   and   KL could be determined from the slope and intercept of linearized equation plot ( C e q e against C e). [50,111]
The dimensionless characteristic of Langmuir isotherm referring to the separation factor (RL) is useful in predicting the adsorption efficiency of the adsorption process. It is an indicator of Langmuir isotherm suitability as either unfavorable (RL > 1), linear (RL = 1), favorable (0 < RL< 1), or irreversible (RL = 0); RL can be express as: [3,112]
R L = 1 1 + K L · C o
where Co is the highest initial concentration of the adsorbate (mg/L), and KL (L/mg) is the Langmuir constant.
To justify the appropriateness of a model that best describes the process, the correlation coefficient (r2) and nonlinear isotherm plots can be used. Previously, it was demonstrated that the Langmuir isotherm can best describe a developed activated carbon monolith/Co3O4 adsorbent in the simultaneous SO2 and NOx removal from flue gas [3]. Based on the study, the Langmuir isotherm suitability was shown by the favorable RL values of 0.3277 and 0.0097, and r2 of 0.9759 and 0.9995 for SO2 and NOx; further evidence was shown by the nonlinear isotherm plots as illustrated in Figure 7. The Langmuir isotherm described the experimental data best according to the alignment of the experimental adsorption capacity (q-experiment) with the q-Langmuir when compared to q-Freundlich. Other studies on Pseudo-first order and Pseudo-second order adsorption kinetic models are reported elsewhere [101,111,113].

4.3. Pressure Drop, Mass Transfer and Heat Transfer in Monolith

Pressure drop is defined as the energy dissipated as a result of fluid flow through a reactor bed. Due to wall friction, orifice effect in the entry region and between the monolith stacks, gas phase acceleration and pressure fall as a result of the gas–liquid distributor, and pressure drops may occur in the monolith [39]. The monolith has the advantages of a large surface area and uniform flow distribution, hence it overcome the pressure-drop concern, which can mitigate investment and operation costs.
Mass transfer studies are rather scarce and incomplete concerning monolith structures [82]. In general, molecules diffuse to the interior of the adsorbent by radial concentration gradient and the molecular diffusion occurs in two ways: [114,115]
  • Pore diffusion occurs as ordinary diffusion when the pore diameter of the adsorbent is large compared to the mean free path of the adsorbed molecule or as Knudsen diffusion when the pores are much smaller than the mean free path of the gas. The diffusion can also occur in the adsorption process as both ordinary and Knudsen diffusions.
  • Surface diffusion: molecules are adsorbed and transported from one site to another along the pore wall in the direction of decreasing concentration, however surface diffusion may be ignored as it contributes little to the overall transportation.
The structural form of a porous catalytic pellet consists of a large number of irregularly shaped pores interconnected [115]. When the catalyst temperature goes higher, the amount adsorbed decreases but for a constant temperature and fixed adsorbate concentration; an equation for the rate of adsorption is given as: [116]
d α d τ = k ( 1 α ) n = A e E / R T ( 1 α ) n
where, k , α, n, A , E , R , T , are rate constant of the adsorption, fractional weight at time τ, reaction order, activation energy, gas constant, absolute temperature respectively. In addition,
α = w f w w f w o
where, w o is initial weight, w is actual weight, and w f final weight, finally, integrating Equation (16) for n = 1 and n ≠ 1 yield Equations (15) and (16):
ln ( 1 α ) = k τ
[ 1 ( 1 α ) ] ( 1 n ) e x p 1 n = k τ
The feed gas flows axially, while mass and energy are transferred through the monolith stream in both axial and radial directions [117]. Accordingly, the convective heat transfer prevails, while conduction is negligible due to point contacts between pellets in pellet catalysts, however conduction heat transfer dominates in the parallel channels of the monolith catalyst as an alternative mechanism by radial and axial heat transport [118,119]. Therefore, radial heat exchange between the monolithic channels is not possible, no convective heat transfer takes place in this direction [113]. An estimated axial heat conductivity is established for the monolith and presented as: [118,120]
k e , a = k s ( 1 ε )
where, k e , a is the effective axial heat conductivity of monolith substrates, k s is the intrinsic thermal conductivity of the support material and ε is the monolith void fraction (or open frontal area, OFA). The radial conductivity k e , r is given as:
k e , r = k s ( 1 ε ) ( 1 + ε )
A highly conductive material enhances radial heat transfer (i.e., monolith structure) and in return reduces thermal runaway risk with the improvement of thermal stability in catalyst [119]. Theoretical analysis showed low heat transfer efficiency of commercial monolithic structures made of corrugated metal sheets due to their structural characteristics depending on the heat transfer properties of the substrate [121]. Yashnik et al. [122] expressed the Nusselt number inside monolith channels as:
h GS = N U C K G d h , c
where, c = 0.571 ( R ec d h , c L ) 2/3.

5. Catalyst Application

The most toxic gases emitted into the atmosphere in the process of fossil fuel combustion are SO2 and NOx [9]. The secondary pollutants from SO2 and NOx include H2SO4, O3, HNO3 and fine particles (PM2.5); they are harmful to humans and to the natural environment [1]. In the field of air pollution control, the recent priority research areas are in the design of desulfurization and denitrification technologies [123]. Figure 8 is showing the global share (%) of fuel for electricity generation.

5.1. SO2 Removal from Flue Gas

SO2 is a colorless and toxic gas that can be oxidized to form sulfuric acid when mixed with water and is one of the main contaminants in the composition of flue gas [124]. The combustion of fossil fuel in manufacturing and power plants accounts for 13.6% and 69.7%, respectively, of global SO2 emissions [10]. Furthermore, SO2 generates aerosols and is the main cause of acid rain which caused considerable damages to properties and the global environment [1]. The convectional technology for SO2 removal from flue gas is the wet flue gas desulfurization (WFGD). Furthermore, the use of liquid–gas reactions and gas–solid reactions with calcites or dolomites as solvents are reported in the literature [125]. Figure 9 shows a typical flue gas desulfurization (FGD) system.
The SO2 adsorption at stack temperature may lead to the low removal rates. Wet FGD is associated with the production of wastewater in coal-fired power plants and the liquid–solid droplets caused by FGD desulfurized flue gas may be deposited on the corrosion samples [127]. The byproducts from the wet FGD process, such as gypsum, ammonia sulfate or magnesium sulfate, are considered as a burden being secondary pollutants [1]. The FGD process is also non-regenerable, so it could be assumed as turning the problem of air pollution into a problem of solid or liquid pollution [14,128,129].

5.2. NOx Removal from Flue Gas

The term NOx is referred to as the oxides of nitrogen (NO, NO2, NO3, N2O3, N2O4, N2O5, NO, and NO2) [10]. NO2 comprises of 95% of NOx produced as a result of combustion processes and is stated to contribute to air emissions by several researchers [81,124,130]. NO2 is a red-brown gas that can readily form nitric acid when it is reacted with H2O. Among others, photochemical smog and formation of acid rain are the problems caused as a result of NO oxidation to NO2 in air [1].
The selective catalytic reduction (SCR) technology is matured and widely favored among other technologies used in the NOx removal process [1,64]. In a typical denitrification process, urea (CO(NH2)2) or ammonia (NH3) is injected into flue gas stream to reacts with NOx at about 350 °C to produce nitrogen (N2) and water [131]. However, at temperatures <350 °C, ammonium sulfates such as NH4SO4 can be formed due to SO2 reaction with NH3 in the presence of H2O and O2, while its accumulation on the catalyst active site can lead to catalyst deactivation [132].
Kasaoka et al. [133] demonstrated the reduction of NOx in flue gas to N2 with NH3 by a dry process at 350 °C and in a wet process where SO was absorbed at around 50 °C with wastewater byproduct. According to the following expressions, the NO removal efficiency can be achieved in the presence of oxygen [134].
4NO + 4NH3 + O2 → 8N + 6H2O
6NO2 + 8NH3 → 7N2 + 12H2O
NO + NO2 + 2NH3 → 2N2 + 3H2O
SO2 reacts with the water in the flue gas and urea to form sulfuric acid.
SO3 + H2O → H2SO4
NH3 + SO3 + H2O → (NH4).HSO4
2NH3 + SO3 + H2O → (NH4)2SO4
The denitrification technique by the SCR system is associated with poor catalyst durability, ammonia slip, low NOx reduction, and a high temperature requirement [10]. Figure 10 is showing conventional selective catalytic reduction technology.
The WFGD and SCR for SO2 and NOx reduction are the most widely used technology for desulfurization/denitrification, however they are expensive [1,136]. It is therefore necessary to develop cost-effective methods that can simultaneously capture the pollutants.

5.3. Simultaneous SO2/NOx Removal from Flue Gas

The requirement of emission standards for SO2 and NOx in thermal power plants recently is rigorous (35 mg/m3) [137]. Dry carbon-based desulfurization does not emit secondary pollutants, with benefits through the production of elemental sulfur, feasible simultaneous SO2, and NOx removal, cheaper investment than wet scrubbing that requires a wastewater treatment system and poor extraction efficiency of dry scrubbing [7,8,114]. Due to high efficiency in operation, the wet process is used more commonly where calcareous or lime-based scrubbing is utilized, but these approaches are non-regenerational with the transformation of air pollution problem into a problem of solid or liquid emissions [8,14,138].
The multistep reduction is particularly complex with high capital cost, solvent losses, undesired foaming, flooding, fouling of equipment, and corrosion [9]. Therefore, it is better to integrate two or more separation processes into a simultaneous removal of SO2, and NOx from flue gas in order to increase efficiency and address the stated concerns.
Ye et al. [44] reported the simultaneous removal of SO2 and NO from flue gases at 350–400 °C with the space velocity (GHSV) of 2800 h−1 and inlet gas containing 1960 ppm SO2 and 500 ppm NO. Previous study showed that 5% NH3 at 400 °C can directly convert SO2 that is regenerated from catalyst into solid ammonium sulfur salts [139]. The removal efficiencies of SO2 and NO, respectively, were defined as: [8]
SO 2 = SO 2 in SO 2 out SO 2 in   100 ,
NO 2 = NO in NO out NO in   100 %
Much of the literature gave detailed descriptions on the issues of SO2 and NOx removal [139,140,141,142,143,144,145,146]. Table 5 shows some experimental work available in literature on the simultaneous NOx/SOx removal studies.

6. Regeneration

The regeneration process eliminates the contaminants that are stored on the surfaces of the sorbents without altering the porosity or causing adsorbent mass losses and restores the adsorptive power. A major problem with the non-regenerative system is the creation of large quantities of waste sludge to be disposed of, and this has become an obstacle. Saturated sorbents are burned or disposed of in landfills, however this choice is an environmental and economic setback. Regenerative processes are carried out to avoid problems associated with the solid waste disposal and to reduce the cost of adsorbent replacement. The spent catalyst may be a source of ultra-fine material to partially substitute cement or be used as a sand replacement [90]. Regeneration is classified into microbiological, chemical, thermal and regeneration with water [140].
The stability of the catalyst and subsequent activity is affected by the regeneration method [7]. Piotrowski et al. [141] found that little information on regeneration kinetics were available. The reversible physical adsorption can be regenerated by reducing the pressure or increasing the adsorbent temperature, while the reduction gas can regenerate an irreversible water–solid reaction mechanism, such as sulfation on a chemical adsorbent [14].
Previous studies have shown that sulfuric deactivated activated carbon can be heated to above 350 °C, and sulfuric acid is reduced to produce concentrated SO2 under inert gases or by heating in a reductive gas flow (NH3, urea, CO, CH4, H2) [7,42]. NH3 is a favorable reductant for Al2(SO4)3 in contrast to H2 and CH4, due to its regeneration potential on the catalyst at the temperatures of the SO2 removal (400 °C) [142].
It is possible to further transform the off-gas regenerator into sulphuric acid or elemental sulphur [10,143]. Desulfation of CuO/γ-Al2O3 adsorbent can be carried out through the Claus process, under reduced flow of H2, CH4, and NH3, while desorbed SO2 is converted into sulfuric acid by further oxidation or elemental sulfur [144]. Regeneration has been found to serve three functions in SO2 and NO removal by NH3; desorption of the adsorbed sulfur particles, surface alteration (chemical effect), and NH3 preservation on the surface (physical effect) that promotes significant removal of NO at lower temperatures [139].
Air as a regenerating agent is cheap and requires simpler regeneration systems than using ozone (O3); however, O3 can remove pollutants at ambient temperature [140], where temperature is reported to be probably the most critical parameter in regeneration [145]. Previously, Jia et al. [146] described the removal of coke from zeolites at low temperatures.
The pretreatment of catalyst before regeneration is recommended with deionized water and sulfuric acid (pH = 2) washing followed by drying at 60 °C for 10 h [149]. Poisoned catalysts can hardly be regenerated; therefore, poisoned content of the feed must be decreased to acceptable levels [79]. The regeneration efficiency is the measurement and comparison of the amounts of adsorbate retained by the regenerated and the original adsorbent after adsorptions under the same conditions.
R E = q reg q orig   100 %
where, q reg and q orig are the adsorption capacities per unit of mass of the regenerated and the original adsorbent, respectively [140].
Dey and Dhal [97] found that the regeneration can restore the catalyst’s active sites and brings the catalyst to its original state or one of even higher activity. The illustration of the regeneration of hopcalite catalysts after CO oxidation is presented in Figure 11.
The reaction activity of hopcalite is presented as:
CO + Mn4+ → CO+ads + Mn3+
The promotion of Cu has further been associated to the reduction of O2:
1/2O2 + Cu1+ → CO2
The oxidation occurs by the process:
O̶ ads + CO+ads → CO2
The resonance reaction system brings the catalyst back to the active state:
Cu1++ Mn3+ ⇄ Cu1+ + Mn4+
The catalyst V/AC was used for SO2 capture, and the experiment was terminated when the SO2 removal efficiency decreased to 80% and the workers performed thermal regeneration at 300–400 °C for 1 h under N2 (400 mL/min) then further conducted CO regeneration at 300–400 °C in a furnace for 1 h under 0.15% CO [7]. The thermal swing adsorption (TSA) is a process where desorption is achieved by increasing the system temperature while decreasing the pressure; the process is called pressure swing adsorption (PSA) [86].
Contaminants are desorbed in the reverse order or countercurrent to which they adsorbed in a co-current method in monolith, the post regeneration activity follows: drying at 60 °C for 2 h and then at 120 °C for 2 h [150]. Several studies demonstrated the regeneration of catalysts [7,41,151,152,153].

Overview on Fluidized-Bed/Fixed-Bed Reactors for Activity Test

When the fluid carrying channel(s) has size below 1 mm, the reactors are regarded as being micro-structured [154], and a good example is the monolith. Experimental data in the laboratory (i.e., kinetics, adsorption capacity) using fluidized or fixed-bed reactors can help in understanding best how the system works [115]. Most of the two mainly used reactor types for flue gas cleaning laboratory research are the fluidized-bed and fixed-bed reactors. Table 6 showed the merits and demerits of each reactor.

7. Conclusions

The development of a catalyst that can clean up large volumes of flue gases with valuable byproducts such as sulfuric acid with environmental consideration through the elimination of the waste materials and proper disposal have gained a lot of attention. Such a catalyst should be compatible and easily retrofitted to the existing power plants, fired boilers, incinerators, and other chemical process industries with more than 90% efficiency regeneration. Moreover, high surface area with abundant active site and chemical nature are stringent characteristics of any catalyst to perform effectively. In this study, the basis of developing a monolithic supported metal-oxide-based catalyst for environmental application was explored. The possibilities of catalyst deactivation, regeneration, and practical environmental application in SOx and NOx removal from flue gas were discussed.

Author Contributions

Conceptualization, K.S. and U.R.; methodology, K.S. and U.R.; validation, K.S.; W.A.W.A.K.G. and U.R.; investigation, K.S. and U.R.; resources, K.S., W.A.W.A.K.G. and U.R.; data curation, K.S.; writing—original draft preparation, K.S. and U.R.; writing—review and editing, K.S., W.A.W.A.K.G., T.S.Y.C., and U.R.; visualization, K.S. and U.R.; supervision, W.A.W.A.K.G., T.S.Y.C., and U.R.; project administration, W.A.W.A.K.G., T.S.Y.C., and U.R.; funding acquisition, W.A.W.A.K.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Universiti Putra Malaysia (UPM) grant number GP-IPB/2016/9515201 and APC was funded by Research Management Center (RMC), Universiti Putra Malaysia (UPM), Malaysia.

Acknowledgments

The authors would like to gratefully acknowledge Universiti Putra Malaysia (UPM) for the financial support of this work (via Geran Putra-IPB, UPM GP-IPB/2016//9515200, GP-IPB/2016/9515201, and GP-IPB/2016/9515202).

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

ACMActivated Carbon Monolith
ACActivated Carbon
CO2Carbon Dioxide
MOMetal Oxide
NOxNitrogen Oxide
qAdsorption Capacity
RERegeneration Efficiency
SCRSelective Catalytic Reduction
SO2Sulfur dioxide
WFGDWet Flue Gas Desulphurization

References

  1. Cheng, G.; Zhang, C. Desulfurization and denitrification technologies of coal-fired flue gas. Pol. J. Environ. Stud. 2018, 272, 481–489. [Google Scholar] [CrossRef]
  2. Koplitz, S.N.; Jacob, D.J.; Sulprizio, M.P.; Myllyvirta, L.; Reid, C. Burden of disease from rising coal-fired power plant emissions in Southeast Asia. Environ. Sci. Technol. 2016, 51, 1–10. [Google Scholar] [CrossRef] [Green Version]
  3. Kiman, S.; Ghani, W.A.W.A.K.; Choong, T.S.Y.; Rashid, U. Breakthrough studies of Co3O4 supported activated carbon monolith for simultaneous SO2/NOx removal from flue gas. Fuel Process. Technol. 2018, 180, 155–165. [Google Scholar]
  4. Lee, D.W.; Yoo, B.R. Metal oxide (supported) catalysts: Synthesis and applications. J. Ind. Eng. Chem. Adv. 2014, 20, 3947–3959. [Google Scholar] [CrossRef]
  5. Rau, J.Y.; Tseng, H.H.; Chiang, B.C.; Wey, M.Y.; Lin, M.D. Evaluation of SO2 oxidation and fly ash filtration by an activated carbon fluidized-bed reactor: The effects of acid modification, copper addition and operating condition. Fuel 2010, 89, 732–742. [Google Scholar] [CrossRef]
  6. Hu, H.; Wang, S.; Zhang, X.; Zhao, Q.; Li, J. Study on simultaneous catalytic reduction of sulfur dioxide and nitric oxide on rare earth mixed compounds. J. Rare Earths 2006, 24, 24695–24698. [Google Scholar] [CrossRef]
  7. Hou, Y.; Zhang, R.; Han, X.; Huang, Z.; Cui, Y. The mechanism of CO regeneration on V2O5/AC catalyst and sulfur recovery. Chem. Eng. J. 2017, 316, 744–750. [Google Scholar] [CrossRef]
  8. Liu, Z.S. Adsorption of SO2 and NO from incineration flue gas onto activated carbon fibers. Waste Manag. 2008, 28, 2329–2335. [Google Scholar] [CrossRef]
  9. Hajari, A.; Atanga, M.; Hartvigsen, J.L.; Rownaghi, A.A.; Rezaei, F. Combined flue gas cleanup process for simultaneous removal of SOx, NOx, and CO2: A techno-economic analysis. Energ. Fuels 2017, 31, 4165–4172. [Google Scholar] [CrossRef]
  10. Liu, Y.; Bisson, T.M.; Yang, H.; Xu, Z. Recent developments in novel sorbents for flue gas clean up. Fuel Process. Technol. 2010, 91, 1175–1197. [Google Scholar] [CrossRef]
  11. Sinha, S.; Chattopadhyay, S. A study on application of renewable energy technologies for mitigatting the adverse environmental impacts generated from power generation units in Himalayan region. Int. J. Innov. Res. Sci. Technol. 2016, 3, 212–232. [Google Scholar]
  12. Hosseini, S.; Marahel, E.; Bayesti, I.; Abbasi, A.; Abdullah Choong, C.T.S.Y. CO2 adsorption on modified carbon coated monolith: Effect of surface modification by using alkaline solutions. Appl. Surf. Sci. 2015, 324, 569–575. [Google Scholar] [CrossRef]
  13. Ma, S.; Zhao, Y.; Yang, J.; Zhang, S.; Zhang, J.; Zheng, C. Research progress of pollutants removal from coal-fired flue gas using non-thermal plasma. Renew. Sustain. Energ. Rev. 2017, 67, 791–810. [Google Scholar] [CrossRef]
  14. Lin, Y.S.; Deng, S.G. Removal of trace sulfur dioxide from gas stream by regenerative sorption processes. Sep. Purif. Technol. 1998, 13, 65–77. [Google Scholar] [CrossRef]
  15. Gupta, A.; Gaur, V.; Verma, N. Breakthrough analysis for adsorption of sulfur-dioxide over zeolites. Chem. Eng. Process. Process Intensif. 2004, 43, 9–22. [Google Scholar] [CrossRef]
  16. Trimm, D.L. The regeneration or disposal of deactivated heterogeneous catalysts. Appl. Catal. A Gen. 2001, 212, 153–160. [Google Scholar] [CrossRef]
  17. Jia, L.Y.; AlFarouha, L.; Pinard, S.; Hedan, J.D.; Comparot, A.; Dufour, K.; Ben, T.H.; Vezin, C.; Batiot, D. Environmental new routes for complete regeneration of coked zeolite. Appl. Catal. B Environ. 2017, 219, 82–91. [Google Scholar] [CrossRef]
  18. Tsybulevski, A.M.; Tkachenko, O.P.; Rode, E.J.; Weston, K.C.; Kustov, L.M.; Sulman, E.M.; Doluda, V.Y.; Greish, A.A. Reactive adsorption of sulfur compounds by transition metal polycation exchanged zeolites and desulfurization of hydrocarbon streams. Energ. Technol. 2017, 5, 1627–1637. [Google Scholar] [CrossRef]
  19. Ju, F.; Liu, C.; Li, K.; Meng, C.; Gao, S.; Ling, H. Reactive adsorption desulfurization of FCC gasoline over a Ca-Doped Ni-ZnO/Al2O3-SiO2. Energ. Fuels 2016, 30, 6688–6697. [Google Scholar] [CrossRef]
  20. Tang, L.; Zhao, Z.; Li, K.; Yu, X.; Wei, Y.; Liu, J.; Peng, Y.; Li, Y.; Chen, Y. Highly active monolith catalysts of LaKCoO3 perovskite-type complex oxide on alumina-washcoateddiesel particulate filter and the catalytic performances for the combustion of soot. Catal. Today 2020, 339, 159–173. [Google Scholar] [CrossRef]
  21. Zhang, L.; Jiang, Y.; Chen, B.; Shi, C.; Li, Y.; Wang, C. Exceptional activity for formaldehyde combustion using siliceous beta zeolite as a catalyst support. Catal. Today 2020, 339, 174–180. [Google Scholar] [CrossRef]
  22. Davó-quiñonero, A.; Sorolla-rosario, D.; Bailón-garcía, E.; Lozano-castelló, D. Improved asymmetrical honeycomb monolith catalyst prepared using a 3D printed template. J. Hazard. Mater. 2019, 368, 638–643. [Google Scholar] [CrossRef] [PubMed]
  23. José, M.; Gatica, J.; Castiglioni, C.; Santos, M.; Pilar, Y.; Gustavo, C.; Martín, T.; Hilario, V. Use of pillared clays in the preparation of washcoatedclay honeycomb monoliths as support of manganese catalysts for the total oxidation of VOCs. Catal. Today 2017, 296, 84–94. [Google Scholar]
  24. Tahay, P.; Khani, Y.; Jabari, M.; Bahadoran, F.; Safari, N. Highly porous monolith/TiO2 supported Cu, Cu-Ni, Ru, and Pt catalysts in methanol steam reforming process for H2 generation. Appl. Catal. A Gen. 2018, 554, 44–53. [Google Scholar] [CrossRef]
  25. Kiman, S.; Ghani, W.A.W.A.K.; Choong, T.S.Y.; Rashid, U. Regeneration/optimization of activated carbon monolith in simultaneous SO2/NOx removal from flu gas. Chem. Eng. Technol. 2019, 9, 1928–1940. [Google Scholar]
  26. Sharma, M.; Choudhury, D.; Hazra, S.; Basu, S. Effective removal of metal ions from aqueous solution by mesoporous MnO2 and TiO2 monoliths: Kinetic and equilibrium modelling. J. Alloys Compd. 2017, 720, 221–229. [Google Scholar] [CrossRef]
  27. Darunte, L.A.; Terada, Y.; Murdock, C.R.; Walton, K.S.; Sholl, D.S.; Jones, C.W. Monolith supported amine functionalized Mg adsorbents for CO2 capture. ACS Appl. Mater. Interfaces 2017, 2, 17042–17050. [Google Scholar] [CrossRef] [Green Version]
  28. Zhu, L.; Tan, C.F.; Gao, M.; Ho, G.W. Design of a metal oxide-organic framework (MoOF) foam microreactor: Solar-induced direct pollutant degradation and hydrogen generation. Adv. Mater. 2015, 27, 7713–7719. [Google Scholar] [CrossRef]
  29. Shen, F.; Liu, J.; Dong, Y.; Wu, D.; Gu, C.; Zhang, Z. Elemental mercury removal from syngas by porous carbon-supported CuCl2 sorbents. Fuel 2019, 239, 138–144. [Google Scholar] [CrossRef]
  30. Abubakar, U.C.; Alhooshani, K.R.; Adamu, S.; Al Thagfi, J.; Saleh, T.A. The effect of calcination temperature on the activity of hydrodesulfurization catalysts supported on mesoporous activated carbon. J. Clean. Prod. 2019, 211, 1567–1575. [Google Scholar] [CrossRef]
  31. Lin, Y.; Li, Y.; Xu, Z.; Xiong, J.; Zhu, T. Transformation of functional groups in the reduction of NO with NH3 over nitrogen-enriched activated carbons. Fuel 2018, 223, 312–323. [Google Scholar] [CrossRef]
  32. Shiung, S.; Rock, L.; Liew, K.; Mun, Y.; Elfina, W. Activated carbon for catalyst support from microwave pyrolysis of orange peel. Waste Biomass Valor. 2017, 8, 2109–2119. [Google Scholar]
  33. Kiman, S.; Ghani, W.A.W.A.K.; Choong, T.S.Y.; Rashid, U. Carbonaceous materials modified catalysts for simultaneous SO2/NOx removal from flue gas: A review. Catal. Rev. Sci. Eng. 2019, 61, 134–161. [Google Scholar]
  34. Patton, A.; Crittenden, B.D.; Perera, S.P. Use of the linear driving force approximation to guide the design of monolithic adsorbents. Chem. Eng. Res. Des. 2004, 82, 999–1009. [Google Scholar] [CrossRef]
  35. Kreutzer, M.T.; Du, P.; Heiszwolf, J.J.; Kapteijn, F.; Moulijn, J.A. Mass transfer characteristics of three-phase monolith reactors. Chem. Eng. Sci. 2001, 56, 6015–6023. [Google Scholar] [CrossRef]
  36. Heiszwolf, J.J.; Engelvaart, L.B.; Van den Eijnden, M.G.; Kreutzer, M.T.; Kapteijn, F.; Moulijn, J.A. Hydrodynamic aspects of the monolith loop reactor. Chem. Eng. Sci. 2001, 56, 805–812. [Google Scholar] [CrossRef]
  37. Moreno-Castilla, C.; Pérez-Cadenas, A.F. Carbon-based honeycomb monoliths for environmental gas-phase applications. Materials 2010, 3, 1203–1227. [Google Scholar] [CrossRef] [Green Version]
  38. Boger, T.; Heibel, A.K.; Sorensen, C.M. Monolithic catalysts for the chemical industry. Ind. Eng. Chem. Res. 2004, 43, 4602–4611. [Google Scholar] [CrossRef]
  39. Roy, S.; Bauer, T.; Al-Dahhan, M.; Lehner, P.; Turek, T. Monoliths as multiphase reactors: A review. AIChE J. 2004, 50, 2918–2938. [Google Scholar] [CrossRef]
  40. Hosseini, S.; Rashid, S.A.; Abbasi, A.; Babadi, F.E.; Abdullah, L.C.; Choong, T.S.Y. Effect of catalyst and substrate on growth characteristics of carbon nanofiber onto honeycomb monolith. Rev. Mex. Urol. 2016, 76, 440–449. [Google Scholar] [CrossRef]
  41. Lei, Z.; Long, A.; Jia, M.; Liu, X. Experimental and kinetic study of selective catalytic reduction of NO with NH3 over CuO/Al2O3/cordierite catalyst. Chin. J. Chem. Eng. 2010, 18, 721–729. [Google Scholar] [CrossRef]
  42. Boyano, A.; Lázaro, M.J.; Cristiani, C.; Maldonado-hodar, F.J.; Forzatti, P.; Moliner, R. A comparative study of V2O5/AC and V2O5/Al2O3 catalysts for the selective catalytic reduction of NO by NH3. Chem. Eng. J. 2009, 149, 173–182. [Google Scholar] [CrossRef]
  43. Campanati, M.; Fornasari, G.; Vaccari, A. Fundamental in the preparation of heterogeneous catalysts. Catal. Today 2003, 77, 299–314. [Google Scholar] [CrossRef]
  44. Liu, Y.; Ning, P.; Li, K.; Tang, L.; Hao, J.; Song, X.; Zhang, G.; Wang, C. Simultaneous removal of NOx and SO2 by low-temperature selective catalytic reduction over modified activated carbon catalysts. Russ. J. Phys. Chem. A 2017, 91, 490–499. [Google Scholar] [CrossRef]
  45. Han, L.; Gao, M.; Hasegawa, J.; Li, S.; Shen, Y.; Li, H.; Shi, L.; Zhang, D. SO2-tolerant selective catalytic reduction of NOx over meso-TiO2@Fe2O3@Al2O3 metal-based monolith catalysts. Environ. Sci. Technol. 2019, 53, 6462–6473. [Google Scholar] [CrossRef]
  46. Mendes, N.; Ozhan, C.; Da Costa, P.; Bacariza, M.C.; Henriques, C. Optimizing washcoatingconditions for the preparation of zeolite-based cordierite monoliths for NOx CH4-SCR: A required step for real application. Ind. Eng. Chem. Res. 2019, 58, 11799–11810. [Google Scholar]
  47. Li, X.; Zhang, X.; Zhong, L.; Zhang, C.; Fang, Q.; Chen, G. Enhancement of SCR performance of monolithic Mn–Ce/Al2O3/cordierite catalysts by using modified deposition precipitation method. Asia-Pac. J. Chem. Eng. 2019, 14, e2318. [Google Scholar] [CrossRef]
  48. Qi, K.; Xie, J.; Li, F.; He, F. Experimental study on preparation and operating conditions over a promising monolithic catalyst for NOx removal: MnOx/TiO2/cordierite. Mater. Sci. Forum 2017, 898, 1905–1915. [Google Scholar] [CrossRef]
  49. Julkapli, N.M.; Bagheri, S. Graphene supported heterogeneous catalysts: An overview. Int. J. Hydrogen Energy 2014, 40, 948–979. [Google Scholar] [CrossRef]
  50. Teoh, Y.P.; Khan, M.A.; Choong, T.S.Y. Kinetic and isotherm studies for lead adsorption from aqueous phase on carbon coated monolith. Chem. Eng. J. 2013, 217, 248–255. [Google Scholar] [CrossRef]
  51. Ma, Z.; Zaera, F. Heterogeneous catalysis by metals. Encycl. Inorg. Bioinorg. Chem. 2011, 1–17. [Google Scholar]
  52. Zhang, F.; Tian, X.; Shah, M.; Yang, W. Synthesis of magnetic carbonaceous acids derived from hydrolysates of Jatropha hulls for catalytic biodiesel production. RSC Adv. 2017, 7, 11403–11413. [Google Scholar] [CrossRef] [Green Version]
  53. Guo, P.; Huang, F.; Zheng, M. Magnetic solid base catalysts for the production of biodiesel. J. Am. Oil Chem. Soc. 2012, 89, 925–933. [Google Scholar] [CrossRef]
  54. Xue, B.; Luo, J.; Zhang, F.; Fang, Z. Biodiesel production from soybean and Jatropha oils by magnetic. Energy 2014, 68, 584–591. [Google Scholar] [CrossRef]
  55. Feyzi, M.; Norouzi, L. Preparation and kinetic study of magnetic Ca/Fe3O4@SiO2 nanocatalysts for biodiesel production. Renew. Energ. 2016, 94, 579–586. [Google Scholar] [CrossRef]
  56. Zhang, F.; Fang, Z.; Wang, Y. Biodiesel production directly from oils with high acid value by magnetic Na2SiO@Fe3O4/C catalyst and ultrasound. Fuel 2015, 150, 370–377. [Google Scholar] [CrossRef]
  57. Fadhil, A.B.; Aziz, A.M.; Al-tamer, M.H. Biodiesel production from Silybummarianum L. seed oil with high FFA content using sulfonated carbon catalyst for esterification and base catalyst for transesterification. Energ. Convers. Manag. 2016, 108, 255–265. [Google Scholar] [CrossRef]
  58. Cao, M.O.; Liu, Y.; Zhang, P.; Fan, M.; Jiang, P. Biodiesel production from soybean oil catalyzed by magnetic. Fuel 2016, 164, 314–321. [Google Scholar]
  59. Stankiewicz, A. Process Intensity, cation in-line monolithic reactor. Chem. Eng. Sci. 2001, 56, 359–364. [Google Scholar] [CrossRef]
  60. Li, J.; Liang, X. Magnetic solid acid catalyst for biodiesel synthesis from waste oil. Energ. Conver. Manag. 2017, 141, 126–132. [Google Scholar] [CrossRef]
  61. Dai, H. Environmental catalysis: A solution for the removal of atmospheric pollutants. Sci. Bull. 2015, 60, 1708–1710. [Google Scholar] [CrossRef] [Green Version]
  62. Ruiz-Martínez, E.; Sánchez-Hervás, J.M.; Otero-Ruiz, J. Effect of operating conditions on the reduction of nitrous oxide by propane over a Fe-Zeolite monolith. Appl. Catal. B Environ. 2005, 61, 306–315. [Google Scholar]
  63. Radwan, N.R.E.; El-Shall, M.S.; Hassan, H.M.A. Synthesis and characterization of nanoparticle Co3O4, CuO and NiOcatalysts prepared by physical and chemical methods to minimize air pollution. Appl. Catal. A Gen. 2007, 331, 8–18. [Google Scholar] [CrossRef]
  64. Huang, Y.; Gao, D.; Tong, Z.; Zhang, J.; Luo, H. Oxidation of NO over cobalt oxide supported on mesoporous silica. J. Nat. Gas Chem. 2009, 18, 421–428. [Google Scholar] [CrossRef]
  65. Yan, Z.; Wang, J.; Zou, R.; Liu, L.; Zhang, Z.; Wang, X. Hydrothermal synthesis of CeO2 nanoparticles on activated carbon with enhanced desulfurization activity. Energ. Fuels 2012, 26, 5879–5886. [Google Scholar] [CrossRef]
  66. Haakana, T.; Kolehmainen, E.; Turunen, I.; Mikkola, J.P.; Salmi, T. The development of monolith reactors: General strategy with a case study. Chem. Eng. Sci. 2004, 59, 5629–5635. [Google Scholar] [CrossRef]
  67. Kiman, S.; Ghani, W.A.W.A.K.; Choong, T.S.Y.; Rashid, U. Activated carbon monolith Co3O4 based catalyst: Synthesis, characterization and adsorption Studies. Environ. Technol. Innov. 2018, 12, 273–285. [Google Scholar]
  68. Liu, Y.; Deng, J.; Xie, S.; Wang, Z.; Dai, H. Catalytic removal of volatile organic compounds using ordered porous transition metal oxide and supported noble metal catalysts. CuihuaXuebao/Chin. J. Catal. 2016, 37, 1193–1205. [Google Scholar] [CrossRef]
  69. Neyestanaki, A.K.; Klingstedt, F.; Salmi, T.; Murzin, D.Y. Deactivation of postcombustioncatalysts. A review. Fuel 2004, 83, 395–408. [Google Scholar] [CrossRef]
  70. Kiman, S.; Ghani, W.A.W.A.K.; Choong, T.S.Y.; Rashid, U. Optimization of activated carbon monolith Co3O4-based catalyst for simultaneous SO2/NOx removal from flue gas using response surface methodology. Combust. Sci. Technol. 2020, 192, 786–803. [Google Scholar]
  71. Boyano, A.; Lombardo, N.; Moliner, R. Vanadium-loaded carbon-based monoliths for the on-board NO reduction: Experimental study of operating conditions. Chem. Eng. J. 2008, 144, 343–351. [Google Scholar] [CrossRef]
  72. Boyano, A.; Herrera, C.; Larrubia, M.A.; Alemany, L.J.; Moliner, R.; Lázaro, M.J. Vanadium loaded carbon-based monoliths for the on-board no reduction: Influence of temperature and period of the oxidation treatment. Chem. Eng. J. 2010, 160, 623–633. [Google Scholar] [CrossRef]
  73. Couck, S.; Lefevere, J.; Mullens, S.; Protasova, L.; Meyen, V.; Desmet, G.; Baron, G.V.; Denayer, J.F. CO2, CH4 and N2 separation with a 3DFD-printed ZSM-5 monolith. Chem. Eng. J. 2017, 308, 719–726. [Google Scholar] [CrossRef]
  74. Kiel, J.H.A.; Edelaar, A.C.S.; Prins, W.; Van Swaaij, W.P.M. Performance of silica-supported copper oxide sorbents for SO2/NOx-removal from flue gas. Appl. Catal. B Environ. 1992, 141–160. [Google Scholar]
  75. Bin, H.; Fang, J.T.; Zhao, Y.T.; Fang, J.J.; Huang, Y. Selective oxidation of hydrogen sulfide to sulfur over activated carbon-supported metal oxides. Fuel 2013, 108, 143–148. [Google Scholar]
  76. Nijhuis, T.A.; Kreutzer, M.T.; Romijn, C.J.; Kapteijn, F.; Moulijn, J.A. Monolithic catalysts as efficient three-phase reactors. Chem. Eng. Sci. 2001, 56, 823–829. [Google Scholar] [CrossRef]
  77. Bartholomew, C.H. Mechanism of catalyst deactivation. Appl. Catal. A Gen. 2001, 212, 17–60. [Google Scholar] [CrossRef]
  78. Bartholomew, C.H.; Argyle, M.D. Advances in catalyst deactivation and regeneration. Catalysts 2015, 5, 949–954. [Google Scholar] [CrossRef]
  79. Worstell, J. Catalyst deactivation. Adiabatic Fixed Bed React. 2014, 52, 35–65. [Google Scholar]
  80. Moulijn, J.A.; Van Diepen, A.E.; Kapteijn, F. Catalyst deactivation: Is it predictable? What to do? Appl. Catal. A Gen. 2001, 212, 3–16. [Google Scholar] [CrossRef]
  81. Kiełtyka, M.; Dias, A.P.S.; Kubiczek, H.; Sarapata, B.; Grzybek, T. The influence of poisoning on the deactivation of Denoxcatalysts. C. R. Chim. 2015, 18, 1036–1048. [Google Scholar] [CrossRef]
  82. Rusu, A.O.; Dumitriu, E. Destruction of volatile organic compounds by catalytic oxidation. Environ. Eng. Manag. J. 2003, 2, 273–302. [Google Scholar] [CrossRef]
  83. Rudyak, V.; Minakov, A. Modeling and optimization of Y-type of micromixer. Micromachines 2009, 33, 75–88. [Google Scholar] [CrossRef]
  84. Marsh, H. Adsorption methods to study microporpsity of coals and carbons. Carbon 1987, 25, 49–58. [Google Scholar] [CrossRef]
  85. Hu, Z.; Srinivasan, M.P.; Ni, Y. Novel activation process for preparing highly microporous and mesoporous activated carbons. Carbon 2001, 39, 877–886. [Google Scholar] [CrossRef]
  86. Hinkov, I.; Deng, J.; Xie, S.; Wang, Z.; Dai, H. Carbon dioxide capture by adsorption. J. Chem. Technol. Metall. 2016, 41, 609–626. [Google Scholar]
  87. Marsh, H.; Rand, B. The characterization of microporous carbons by means of the Dubinin-Radushkevichequation. J. Colloid Interface Sci. 1970, 33, 101–116. [Google Scholar] [CrossRef]
  88. Heigl, N.; Greider, A.; Petter, C.H.; Kolomiets, O.; Sieser, H.W.; Ulbricht, M.; Bonn, G.K.; Huck, C.W. Simultaneous determination of the micro-, meso-, and macropore size fractions of porous polymers by a combined use of fouriertransform hear-infrared diffuse reflection spectroscopy and multivariate techniques. Anal. Chem. 2008, 80, 8493–8500. [Google Scholar] [CrossRef]
  89. Choma, J.; Górka, J.; Jaroniec, M.; Zawislak, A. Development of microporosity in mesoporous carbons. Top. Catal. 2010, 53, 283–290. [Google Scholar] [CrossRef]
  90. Marafi, M.; Stanislaus, A. Spent catalyst waste management:A review. Part I-Developments in hydroprocessingcatalyst waste reduction and use. Resour. Conserv. Recycl. 2008, 52, 859–873. [Google Scholar] [CrossRef]
  91. Lázaro, M.J.; Gálvez, M.T.; Izquierdo, E.; García-Bordejé, C.; Ruiz, R.; Juan, R.; Moliner, J. Novel carbon-based catalysts for the reduction of NO: Influence of support precursors and active phase loading. Catal. Today 2008, 137, 215–221. [Google Scholar] [CrossRef]
  92. Wu, C.; Liang, Y.; Yang, K.; Min, Y.; Liang, Z.; Zhang, L.; Zhang, Y. Clickable periodic mesoporous organosilica monolith for highly efficient capillary chromatographic separation. Anal. Chem. 2016, 88, 1521–1525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Hu, P.; Long, M.; Bai, X.; Wang, C.; Cai, C.; Fu, J. Monolithic cobalt-doped carbon aerogel for efficient catalytic activation of peroxymonosulfate in water. J. Hazard. Mater. 2017, 332, 195–204. [Google Scholar] [CrossRef] [PubMed]
  94. Zhan, S.; Zhang, H.; Zhang, Y.; Shi, Q.; Li, Y.; Li, X.J. Efficient NH3-SCR removal of NOx with highly ordered mesoporous WO3(χ)-CeO2 at low temperatures. Appl. Catal. B Environ. 2017, 203, 199–209. [Google Scholar] [CrossRef] [PubMed]
  95. Rodriguez-reinoso, F. The role of carbon materials in heterogeneouscatalysis. Carbon 1998, 36, 159–175. [Google Scholar] [CrossRef]
  96. Thiruvenkatachari, R.; Su, S.; An, H.; Yu, X.X. Post combustion CO2 capture by carbon fibremonolithic adsorbents. Prog. Energ. Combust. Sci. 2009, 35, 438–455. [Google Scholar] [CrossRef]
  97. Dey, S.; Dhal, G.C. Deactivation and regeneration of Hopcalitecatalyst for carbon monoxide oxidation: A review. Mater. Today Chem. 2019, 14, 100180. [Google Scholar] [CrossRef]
  98. Zhou, X.; Yi, H.; Tang, X.; Deng, H.; Liu, H. Thermodynamics for the adsorption of SO2, NO and CO2 from flue gas on activated carbon fiber. Chem. Eng. J. 2012, 200–202, 399–404. [Google Scholar] [CrossRef]
  99. Sircar, S.; Myers, A.L. Gas adsorption operations: Equilibrium, kinetics, column dynamics and design. Adsorpt. Sci. Technol. 1985, 2, 69–87. [Google Scholar] [CrossRef]
  100. Dabrowski, A. Adsorption from theory to practice. Adv. Colloid Interface Sci. 2001, 93, 135–224. [Google Scholar] [CrossRef]
  101. Singh, V.K.; Kumar, E.A. Comparative studies on CO2 adsorption kinetics by solid adsorbents. Energ. Procedia 2015, 90, 316–325. [Google Scholar] [CrossRef]
  102. Rezaei, F.; Webley, P. Structured adsorbents in gas separation processes. Sep. Purif. Technol. 2010, 70, 243–256. [Google Scholar] [CrossRef] [Green Version]
  103. Podkościelny, P.; Nieszporek, K. Adsorption of phenols from aqueous solutions: Equilibria, calorimetry and kinetics of adsorption. J. Colloid Interface Sci. 2011, 354, 282–291. [Google Scholar] [CrossRef] [PubMed]
  104. Park, K.H.; Balathanigaimani, M.S.; Shim, W.G.; Lee, J.W.; Moon, H. Adsorption characteristics of phenol on novel corn grain-based activated carbons. Microp. Mesop. Mater. 2010, 127, 1–8. [Google Scholar] [CrossRef]
  105. Kalavathy, H.; Regupathi, M.I.; Pillai, M.G.; Miranda, L.R. Modelling, analysis and optimization of adsorption parameters for H3PO4 activated rubber wood sawdust using response surface methodology (RSM). Colloids Surf. B Biointerfaces 2009, 70, 35–45. [Google Scholar] [CrossRef]
  106. Chaudhary, N.; Balomajumder, C. Optimization study of adsorption parameters for removal of phenol on aluminum impregnated fly ash using response surface methodology. J. Taiwan Inst. Chem. Eng. 2014, 45, 852–859. [Google Scholar] [CrossRef]
  107. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Kenneth, S.W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC technical report). Pure Appl. Chem. 2015, 87, 1–19. [Google Scholar] [CrossRef] [Green Version]
  108. Ravennia, G.; Elhamib, O.H.; Ahrenfeldta, J.; Henriksena, U.B.; Neubauerb, Y. Adsorption and decomposition of tar model compounds over the surface of gasification char and active carbon within the temperature range 250–800 °C. Appl. Energ. 2019, 241, 139–151. [Google Scholar] [CrossRef]
  109. Sumathi, S.; Bhatia, S.; Lee, K.T.; Mohamed, A.R. Performance of shell activated carbon impregnated with CeO2 and V2O5 catalyst in simultaneous removal of SO2 and NO. J. Appl. Sci. 2010, 12, 1052–1059. [Google Scholar]
  110. Sircar, S. Comments on practical use of Langmuir gas adsorption isotherm model. Adsorption 2017, 23, 121–130. [Google Scholar] [CrossRef]
  111. Ghorbani, F.; Younesi, H.; Ghasempouri, S.M.; Zinatizadeh, A.A.; Amini, M.; Daneshi, A. Application of response surface methodology for optimization of cadmium biosorption in an aqueous solution by Saccharomyces Cerevisiae. Chem. Eng. J. 2008, 145, 267–275. [Google Scholar] [CrossRef]
  112. Abdedayem, A.; Guiza, M.; Ouederni, A. Copper supported on porous activated carbon obtained by wetness impregnation: Effect of preparation conditions on the ozonation catalyst’s characteristics. C. R. Chim. 2015, 18, 100–109. [Google Scholar] [CrossRef]
  113. Boger, T.; Heibel, A.K. Heat transfer in conductive monolith structures. Chem. Eng. Sci. 2005, 60, 1823–1835. [Google Scholar] [CrossRef]
  114. Lua, A.C.; Guo, J. Adsorption of sulfur dioxide on activated carbon from oil-palm waste. J. Environ. Eng. 2001, 127, 895–901. [Google Scholar] [CrossRef]
  115. Andrigo, P.; Bagatin, R.; Pagani, G.; Hugo, A. Fixed bed reactors. Catal. Today 2011, 52, 197–221. [Google Scholar] [CrossRef]
  116. Yavuz, R. Sulfur dioxide adsorption by activated carbons having different textural and chemical properties. Fuel 2008, 87, 3207–3215. [Google Scholar]
  117. Gu, T.; Balakotaiah, V. Impact of heat and mass dispersion and thermal effects on the scale-up of monolith reactors. Chem. Eng. J. 2016, 284, 513–535. [Google Scholar] [CrossRef] [Green Version]
  118. Visconti, C.G.; Groppi, G.; Tronconi, E. Accurate prediction of the effective radial conductivity of highly conductive honeycomb monoliths with square channels. Chem. Eng. J. 2013, 223, 224–230. [Google Scholar] [CrossRef]
  119. Groppi, G.; Tronconi, E. Honeycomb supports with high thermal conductivity for gas/solid chemical processes. Catal. Today 2005, 105, 297–304. [Google Scholar] [CrossRef]
  120. Tronconi, E.; Groppi, G.; Visconti, C.G. Structured catalysts for non-adiabatic applications. Curr. Opin. Chem. Eng. 2014, 5, 55–67. [Google Scholar] [CrossRef]
  121. Groppi, G.; Tronconi, E.; Cortelli, C.; Leanza, R. Conductive monolithic catalysts: Development and industrial pilot tests for the oxidation of O-xylene to phthalic anhydride. Ind. Eng. Chem. Res. 2012, 51, 7590–7596. [Google Scholar] [CrossRef]
  122. Yashnik, S.A.; Ismagilov, Z.R.; Koptyug, I.V.; Andrievskaya, I.P. Formation of textural and mechanical properties of extruded ceramic honeycomb monoliths: An 1H-NMR imaging study. Catal. Today 2005, 105, 507–515. [Google Scholar] [CrossRef]
  123. Liu, Y.X.; Ziyang, L.; Yan, W.; Yanshan, Y.; Jian, F.P.; Jun, Z.; Qian, W. Simultaneous absorption of SO2 and NO from flue gas using ultrasound/Fe2+/heat coactivated persulfate system. J. Hazard. Mater. 2018, 342, 326–334. [Google Scholar] [CrossRef] [PubMed]
  124. Sun, Y.; Zwolińska, E.; Chmielewski, A.G. Abatement technologies for high concentrations of NOx and SO2 removal from exhaust gases: A review. Crit. Rev. Environ. Sci. Technol. 2016, 46, 119–142. [Google Scholar] [CrossRef]
  125. Jing, W.; Hou, Y.; Guo, Q.; Huang, Z.; Han, X.; Ma, G. Using vanadyl sulfate to prepare carbon-supported vanadium catalyst for flue gas desulfurization. J. Fuel Chem. Technol. 2013, 41, 1223–1233. [Google Scholar] [CrossRef]
  126. Pan, P.; Chen, H.; Liang, Z.; Zhao, Q. Deposition and corrosion characteristics of liquid-solid droplets on tubular corrosion probes in desulfurized flue gas. Eng. Fail. Anal. 2018, 90, 129–140. [Google Scholar] [CrossRef]
  127. Gingerich, D.B.; Grol, E.; Mauter, M.S. Fundamental challenges and engineering opportunities in flue gas desulfurization wastewater treatment at coal fired power. Environ. Sci. Water Res. Technol. 2018, 4, 909–925. [Google Scholar] [CrossRef]
  128. Liu, X.; Khinast, J.G.; Glasser, B.J. A parametric investigation of impregnation and drying of supported catalysts. Chem. Eng. Sci. 2008, 63, 4517–4530. [Google Scholar] [CrossRef]
  129. Yang, Y.; Chiang, K.; Burke, N. Porous carbon-supported catalysts for energy and environmental applications: A short review. Catal. Today 2011, 178, 197–205. [Google Scholar] [CrossRef]
  130. Skalska, K.; Miller, J.S.; Ledakowicz, S. Trends in NOx abatement: A review. Sci. Total Environ. 2010, 408, 3976–3989. [Google Scholar] [CrossRef]
  131. Nova, I.; Beretta, A.; Groppi, G.; Lietti, L.; Tronconi, E.; Forzatti, P. Monolithic catalysts for NOx removal from stationary sources. Struct. Catal. React. 2005, 171–214. [Google Scholar]
  132. Liu, Q.; Liu, Z. Carbon supported vanadiafor multi-pollutants removal from flue gas. Fuel 2013, 108, 149–158. [Google Scholar] [CrossRef]
  133. Kasaoka, S.; Sasaoka, E.; Iwasaki, H. Vanadium oxides (V2Ox) catalysts for dry-type and simultaneous removal of sulfur oxides and nitrogen oxides with ammonia at low temperature. Bull. Chem. Soc. Jpn. 1989, 62, 1226–1232. [Google Scholar] [CrossRef] [Green Version]
  134. Athappan, A.; Sattler, M.L.; Sethupathi, S. Selective catalytic reduction of nitric oxide over cerium-doped activated carbons. J. Environ. Chem. Eng. 2005, 3, 2502–2513. [Google Scholar] [CrossRef] [Green Version]
  135. Wang, C.; Feng, Y.; Mingyuan, Z.Y.; Jianming, D.; Jinli, Z.; Peng, C.; Bin, D. Micro spherical MnO2-CeO2-Al2O3 mixed oxide for monolithic honeycomb catalyst and application in selective catalytic reduction of NOx with NH3 at 50–150 °C. Chem. Eng. J. 2018, 346, 182–192. [Google Scholar] [CrossRef]
  136. Abdullah, A.H.; Mat, R.; Somderam, S.; Aziz, A.S.A.; Mohamed, A. Hydrogen sulfide adsorption by zinc oxide-impregnated zeolite (synthesized from Malaysian Kaolin) for biogas desulfurization. J. Ind. Eng. Chem. 2018, 65, 334–342. [Google Scholar] [CrossRef]
  137. Hao, R.; Yaoyu, Z.; Zhaoyue, W.; Yuanpeng, L.; Bo, Y.; Xingzhou, M.; Yi, Z. An advanced wet method for simultaneous removal of SO2 and NO from coal-fired flue gas by utilizing a complex absorbent. Chem. Eng. J. 2017, 307, 562–571. [Google Scholar] [CrossRef]
  138. Yang, B.; Chen, L.; Sun, F. Recovery of sulfur dioxide from gas mixture in packed bed column. Int. J. Energ. Environ. 2011, 2, 211–218. [Google Scholar]
  139. Liu, Q.; Liu, Z.; Wu, W. Effect of V2O5additive on simultaneous SO2 and NO removal from flue gas over a monolithic cordierite-based CuO/Al2O3 catalyst. J. Fuel Chem. Technol. 2009, 35, 285–289. [Google Scholar]
  140. Salvador, F.; Martin-Sanchez, N.; Sanchez-Hernandez, R.; Sanchez-Montero, M.J.; Izquierdo, C. Regeneration of carbonaceous adsorbents. Part II: Chemical, microbiological and vacuum regeneration. Microp. Mesopor. Mater. 2015, 202, 277–296. [Google Scholar] [CrossRef]
  141. Piotrowski, K.; Wiltowski, T.; Szyma, T.; Mondal, K. Cycling effects on the methane regeneration kinetics of CuO/Al2O3 sorbent. Chem. Eng. J. 2005, 108, 227–237. [Google Scholar] [CrossRef]
  142. Jia, Z.; Liu, Z.Y.; Zhao, Y. Kinetics of SO2 removal from flue gas on CuO/Al2O3 sorbent catalyst. Chem. Eng. Technol. 2007, 30, 1221–1227. [Google Scholar] [CrossRef]
  143. Edelaar, A.C.S.; Prms, W.; Van, W.P.M. Performance of silica-supported copper oxide sorbents for SO2/NO, Removal from flue gas II. Selective catalytic reduction of nitric oxide by ammonia. App. Catal. B Environ. 1992, 1, 41–60. [Google Scholar]
  144. Mathieu, Y.; Tzanis, L.; Soulard, M.; Patarin, J.; Vierling, M.; Molière, M. Adsorption of SOx by oxide materials: A review. Fuel Process. Technol. 2013, 114, 81–100. [Google Scholar] [CrossRef]
  145. Afonso, J.C.; Aranda, D.A.G.; Schmal, M.; Frety, R. Regeneration of a Pt-Sri/Al2O3 catalyst: Influence of heating rate, temperature and time of Regeneration. Fuel Process. Technol. 1997, 50, 35–48. [Google Scholar] [CrossRef]
  146. Jia, Y.; Jiang, J.; Sun, K.; Chen, C. Oxidation of formic acid over palladium catalyst supported on activated carbon derived from polyaniline and modified lignosulfonate composite. J. Fuel Chem. Technol. 2017, 45, 100–105. [Google Scholar] [CrossRef]
  147. Schobing, J.; Tschamber, V.; Brilhac, J.F.; Auclaire, A.; Hohl, Y. Simultaneous soot combustion and NOx reduction over a vanadia-based selective catalytic reduction catalyst. C. R. Chim. 2018, 21, 221–231. [Google Scholar] [CrossRef]
  148. Liu, W.; He, X.; Pang, S.; Zhang, Y. Effect of relative humidity on O3 and NO2 oxidation of SO2 on α-Al2O3 particles. Atmos. Environ. 2017, 167, 245–253. [Google Scholar] [CrossRef]
  149. Shang, X.; Hu, G.; He, C.; Zhao, J.; Zhang, F.; Xu, Y.; Zhang, Y.; Li, J.; Chen, J. Regeneration of full-scale commercial honeycomb monolith catalyst (V2O5-WO3/TiO2) used in coal-fired power plant. J. Ind. Eng. Chem. 2012, 18, 513–519. [Google Scholar] [CrossRef]
  150. Yu, Y.; Meng, X.; Chen, J.; Yin, L.; Qiu, T.; He, C. Deactivation mechanism and feasible regeneration approaches for the used commercial NH3-SCR catalysts. Environ. Technol. 2016, 37, 828–836. [Google Scholar] [CrossRef]
  151. Federica, R.; Michela, A.; Gargiulo, R.; Paola, A. Kinetic study and breakthrough analysis of the hybrid physical/chemical CO2 adsorption/desorption behavior of a magnetite-based sorbent. Chem. Eng. J. 2019, 372, 526–535. [Google Scholar]
  152. Rodrigues, C.P.; Kraleva, V.; Ehrich, H.; Noronha, F.B. Structured reactors as an alternative to fixed-bed reactors: Influence of catalyst preparation methodology on the partial oxidation of ethanol. Catal. Today 2016, 273, 12–24. [Google Scholar] [CrossRef]
  153. Zafer, S.; Oana, M.; Merve, K.; Louise, O.; Emrah, O. Trade-off between NOx storage capacity and sulfur tolerance on Al2O3/ZrO2/TiO2–based DeNOx catalysts. Catal. Today 2019, 320, 152–164. [Google Scholar]
  154. Kolb, G.; Hessel, V. Micro-structured reactors for gas phase reactions. Chem. Eng. J. 2004, 98, 1–38. [Google Scholar] [CrossRef]
  155. Lin, H.; Gao, X.; Luo, Z.; Cen, K.; Huang, Z. Removal of NO. Fuel 2004, 83, 1349–1355. [Google Scholar] [CrossRef]
  156. Yeh, J.T.; Demski, R.J.; Strakey, J.P.; Joubert, J.I. Combined SO2/NOx removal from flue gas. Environ. Prog. 1985, 4, 223–228. [Google Scholar] [CrossRef]
  157. Warnecke, R. Gasification of biomass: Comparison of fixed bed and fluidized bed gasifier. Biomass Bioenergy 2000, 18, 489–497. [Google Scholar] [CrossRef]
Figure 1. Monolith structure.
Figure 1. Monolith structure.
Catalysts 10 01018 g001
Figure 2. SEM/EDX spectra of monolith Co3O4 supported adsorbents. Reproduced with permission from Silas et al. [67]. Copyright 2018 Elsevier Ltd. [67].
Figure 2. SEM/EDX spectra of monolith Co3O4 supported adsorbents. Reproduced with permission from Silas et al. [67]. Copyright 2018 Elsevier Ltd. [67].
Catalysts 10 01018 g002
Figure 3. Schematic representation of synthesis procedure.
Figure 3. Schematic representation of synthesis procedure.
Catalysts 10 01018 g003
Figure 4. Schematic of the catalyst loading methods.
Figure 4. Schematic of the catalyst loading methods.
Catalysts 10 01018 g004
Figure 5. Positions of micro, meso, and macropores on a catalyst active site.
Figure 5. Positions of micro, meso, and macropores on a catalyst active site.
Catalysts 10 01018 g005
Figure 6. Mechanism of CO oxidation over a CuMnOx catalyst. Reproduced with permission from Dey and Dhal [97]. Copyright 2019 Elsevier Ltd. [97].
Figure 6. Mechanism of CO oxidation over a CuMnOx catalyst. Reproduced with permission from Dey and Dhal [97]. Copyright 2019 Elsevier Ltd. [97].
Catalysts 10 01018 g006
Figure 7. Experimental vs Langmuir and Freundlich isotherms (a) SO2 (b) NOx. Reproduced with permission from Silas et al. [3]. Copyright 2018 Elsevier Ltd. [3].
Figure 7. Experimental vs Langmuir and Freundlich isotherms (a) SO2 (b) NOx. Reproduced with permission from Silas et al. [3]. Copyright 2018 Elsevier Ltd. [3].
Catalysts 10 01018 g007
Figure 8. Global share (%) of fuel for electricity generation. Reproduced with permission from Sinha & Chattopadhyay [11]. Copyright 2016 International Journal for Innovative Research in Science & Technology (IJIRST) [11].
Figure 8. Global share (%) of fuel for electricity generation. Reproduced with permission from Sinha & Chattopadhyay [11]. Copyright 2016 International Journal for Innovative Research in Science & Technology (IJIRST) [11].
Catalysts 10 01018 g008
Figure 9. The back end of a coal-fired heating plant. Reproduced with permission from Pan et al. [126]. Copyright 2018 Elsevier Ltd. [126].
Figure 9. The back end of a coal-fired heating plant. Reproduced with permission from Pan et al. [126]. Copyright 2018 Elsevier Ltd. [126].
Catalysts 10 01018 g009
Figure 10. Conventional selective catalytic reduction technology. Reproduced with permission from Wang et al. [135]. Copyright 2018 Elsevier Ltd. [135].
Figure 10. Conventional selective catalytic reduction technology. Reproduced with permission from Wang et al. [135]. Copyright 2018 Elsevier Ltd. [135].
Catalysts 10 01018 g010
Figure 11. Regeneration of hopcalite catalysts after CO oxidation. Reproduced with permission from Dey and Dhal [97]. Copyright 2019 Elsevier Ltd. [97].
Figure 11. Regeneration of hopcalite catalysts after CO oxidation. Reproduced with permission from Dey and Dhal [97]. Copyright 2019 Elsevier Ltd. [97].
Catalysts 10 01018 g011
Table 1. Specification of bare monolith.
Table 1. Specification of bare monolith.
MonolithCellsChemical Compositions
Cross section circularChannel squareSiO2 50.9 ± 1.0%
Surface area 1 cm2/gWall thickness 0.25 ± 0.02 mmAl2O3 35.2 ± 1.0%
Length 2.50 ± 0.02 mmWidth 1.02 ± 0.02 mmMgO 13.9 ± 0.5%
Diameter 2.50 ± 0.02 mmCells 400 (cpsi)Others < 1%
Table 2. The recent application of monoliths as catalysts.
Table 2. The recent application of monoliths as catalysts.
Adsorbent/sCatalyst/s SupportedMethod of DopingApplicationRef.
MonolithCo3O4Pore impregnation,
deposition
precipitation,
hydrothermal methods
Simultaneous SO2/NOx removal from flue gas[3]
MonolithTiO2, Fe2O3One-pot self-assembly methodLow-temperature SO2-tolerant monolithic selective catalytic reduction SCR catalysts with high N2 selectivity[45]
Zeolite-based cordierite monolithsPd−Ce−Zeolite powder catalystImpregnationNOx CH4 selective catalytic reduction[46]
CordieriteMn–Ce/Al2O3Impregnation,
deposition
precipitation,
modified deposition
precipitation,
SO2 and NO conversion[47]
CordieriteMnOx/TiO2Sol-gel-impregnation method,Low-temperature selective catalytic reduction (SCR) of NOx with NH3[48]
Table 3. Reviewed catalyst loading on supports.
Table 3. Reviewed catalyst loading on supports.
AdsorbentCatalystCatalyst LoadingSynthesis MethodApplicationRef.
ACFe(NO3)3.9H2O
Mn(NO3)2
Co(NO3)2.6H2O Ce(NO3)3.6H2O
(V,NH4VO3H2C2O4)
Cu(NO3)2.3H2O
1.79 × 10−2
1.82 × 10−2
1.70 × 10−2
7.3 × 10−3
1.96 × 10−2
1.57 × 10−2
Mmol
Pore volume wetness impregnationOxidation of H2S to elemental sulfur.[75]
MonolithNi(NO3)21 wt%Deposition
precipitation
Comparison of monolith and a trickle-bed catalyst system[76]
ACV2O51 wt%.Pore volume impregnationDesulfurization and regeneration[7]
CordieriteCu/Cu(NO3)2.3H2O1.5%Incipient impregnationModelling of monolithic SCR reactor[41]
Table 4. Types and causes of catalyst deactivation.
Table 4. Types and causes of catalyst deactivation.
Deactivation TypeCause of DeactivationPrevention/TreatmentRef.
Chemical poisoningSulfur, chlorine, heavy metals, halogens, silicones, phosphorus, acid catalysts are poisoned by basic materials, oxide catalysts are poisoned by Pb, Hg, As, Cd.Metal oxide-based catalysts exhibit good resistant to deactivation by poisoning, oxidation can reduce the effect of some poisons/regeneration.[16,79,83]
Thermal sinteringCatalyst exposure to high temperature, high partial pressure of water, crystallite growth on the catalytic phase resulting in loss of catalytic surface area, calcinations process, reduction (fresh or passivated catalyst), reaction (hot spots, maldistribution), or regeneration.Stabilizers are used to fill vacancies in the lattice, employing chemical vapor deposition (CVD)-type treatment to restore the active sites/it is better to avoid sintering from occurring.[16,77,80,83]
FoulingCarbon deposit, ash, soot, rust, and scaleRegeneration/rejuvenation[16,80]
Table 5. Previous works on simultaneous SO2/NOx removal.
Table 5. Previous works on simultaneous SO2/NOx removal.
Objective/CommentCatalytic ActivityRef.
Simultaneous removal of NO and SO2 from flue gas catalysts/uses simulated flue gas.0.8 g of catalyst was used with reaction gas flow rate of 500 cm3 min−1 corresponding to gas hourly space velocity (GHSV) of 30,000 h−1. N2-based gas mixture containing 500 ppm NO, 600 ppm NH3, and 1000 ppm SO2 and 0–5% O2.[44]
Vanadia based SCR catalyst study/uses simulated flue gas.The reactive gas flow, containing 1000 ppm of NOx (100 ppm NO2 and 900 ppm NO), 1000 ppm of NH3, 15% of O2, and 8% of H2O at flow rate of GHSV = 55,000 h−1. [147]
Ammonium sulfate salt deactivation in SCR study/uses simulated flue gas. A simulated flue gas of 620 ppm NO, 620 ppm NH3, 3 vol% H2O, 5.5 vol% O2, and balance Ar.[54]
Simultaneous removal of SO2/NO2 from simulated flue gas.Simulated flue gas of 2000 ppm for SO2, 200 ppm for NO2, and about 5% for O2. The flow rate was controlled at 0.15 m3h−1 with a rotameter.[148]
Table 6. Comparison of the fluidized-bed reactors and fixed-bed reactors.
Table 6. Comparison of the fluidized-bed reactors and fixed-bed reactors.
Fluidized-Bed Reactors (+Merit, -Demerit)Fixed-Bed Reactors (+Merit, -Demerit)Ref.
(+) Fly ash passes through the reactor without plugging(-) Plugging can occur (monolith adsorbent/catalyst reactor remedy)[5,80,155,156,157]
(-) Higher pressure drop(+) Simple and robust construction
(+) Shutdown for replacement of catalyst is not needed.(-) Process must be shut down for reloading the reactor with fresh catalyst
(-) Possible attrition of the bed material(+) Pressure drop is low
(+) Continuous operation and regeneration(-) Bad temperature distribution, Low specific capacity
(+) Lots of processes for different applications in operation
(+) High thermal/mass transfer coefficient, high gas/solid, and solid/solid contact areas are continuous(-) Long period to heat-up
(+) No valve problems(-) Valves are required to isolate the regenerator’s absorber
(+) Less space requirement because of great scale-up(+) Can operate at partial load (20 ± 110%)

Share and Cite

MDPI and ACS Style

Silas, K.; Wan Ab Karim Ghani, W.A.; Choong, T.S.Y.; Rashid, U. Monolith Metal-Oxide-Supported Catalysts: Sorbent for Environmental Application. Catalysts 2020, 10, 1018. https://doi.org/10.3390/catal10091018

AMA Style

Silas K, Wan Ab Karim Ghani WA, Choong TSY, Rashid U. Monolith Metal-Oxide-Supported Catalysts: Sorbent for Environmental Application. Catalysts. 2020; 10(9):1018. https://doi.org/10.3390/catal10091018

Chicago/Turabian Style

Silas, Kiman, Wan Azlina Wan Ab Karim Ghani, Thomas Shean Yaw Choong, and Umer Rashid. 2020. "Monolith Metal-Oxide-Supported Catalysts: Sorbent for Environmental Application" Catalysts 10, no. 9: 1018. https://doi.org/10.3390/catal10091018

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