Next Article in Journal
Digital Twins for Continuous mRNA Production
Next Article in Special Issue
Effects of Different Biodiesel-Diesel Blend Fuel on Combustion and Emission Characteristics of a Diesel Engine
Previous Article in Journal
Brownian Motion and Thermophoretic Effects in Mini Channels with Various Heights
Previous Article in Special Issue
Investigation of the Performance and Emission Characteristics of a Diesel Engine with Different Diesel–Methanol Dual-Fuel Ratios
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Investigation on the Intraphase and Interphase Mass Transfer Limitations for NH3-SCR over Cu-ZSM-5

1
School of Marine Engineering, Jiangsu Shipping College, Nantong 226010, China
2
Merchant Marine College, Shanghai Maritime University, Shanghai 201306, China
*
Author to whom correspondence should be addressed.
Processes 2021, 9(11), 1966; https://doi.org/10.3390/pr9111966
Submission received: 26 September 2021 / Revised: 29 October 2021 / Accepted: 30 October 2021 / Published: 4 November 2021
(This article belongs to the Special Issue Clean Combustion and Emission in Vehicle Power System)

Abstract

:
A systematic modeling approach was scrutinized to develop a kinetic model and a novel monolith channel geometry was designed for NH3 selective catalytic reduction (NH3-SCR) over Cu-ZSM-5. The redox characteristic of Cu-based catalysts and the variations of NH3, NOx concentration, and NOx conversion along the axis in porous media channels were studied. The relative pressure drop in different channels, the variations of NH3 and NOx conversion efficiency were analyzed. The model mainly considers NH3 adsorption and desorption, NH3 oxidation, NO oxidation, and NOx reduction. The results showed that the model could accurately predict the NH3-SCR reaction. In addition, it was found that the Cu-based zeolite catalyst had poor low-temperature catalytic performance and good high-temperature activity. Moreover, the catalytic reaction of NH3-SCR was mainly concentrated in the upper part of the reactor. In addition, the hexagonal channel could effectively improve the diffusion rate of gas reactants to the catalyst wall, reduce the pressure drop and improve the catalytic conversion efficiencies of NH3 and NOx.

1. Introduction

A diesel engine is widely used in ships and automobiles [1] because of its high thermal efficiency [2], low fuel consumption [3], low emission, and high efficiency [4]. However, the nitrogen oxides (NOx) and particulate matter (PM) emitted by diesel engines pose a significant threat to human health and the environment [5]. PM is produced by incomplete combustion at low temperature [6], while NOx is produced by combustion at high temperature and oxygen enrichment [7]. The NOx limits for Euro Ⅴ and Euro Ⅵ are 0.18 g/km and 0.08 g/km, respectively [8]. In order to meet this lower limit, when NOx emission exceeds 0.08 g/km, heavy vehicles need to be equipped with aftertreatment devices, such as selective catalytic reduction (SCR) and Lean NOx Trap (LNT). Nowadays, the world environment is becoming worse and worse. The governments and organizations have formulated relevant emission limitation regulations, which are becoming increasingly stringent [9]. At the same time, the International Maritime Organization (IMO) also developed the marine diesel engine MAPPOL 73/78 anti-pollution convention in the by-law VI of the marine diesel engine NOx emissions to make a strict limit, the by-law and in January 2000 came into force [10]. The 2019 Annual Report on Motor Vehicle Pollution Prevention and Control in China showed that diesel vehicles accounted for a large share of pollution in 2018, with NOx emissions accounting for 70% of total emissions. Facing the problem of global environmental degradation, how to effectively reduce diesel engine emissions has become the main task of sustainable development of human society [11].
Reducing NOx and particulate emissions has become the focus [12]. The SCR has been widely promoted in recent years due to its good technical economy, high reliability, and insensitive to sulfur [13]. However, in the stage of National IV and above emission regulations, NH3-SCR is one of the most commonly used aftertreatment technologies for reducing NOx emission of heavy-duty diesel engine [14]. This technology generates ammonia through chemical reactions such as pyrolysis and hydrolysis of urea. Ammonia is adsorbed and stored on the catalyst surface and then reacts with NOx in exhaust gas to generate nitrogen (N2) and water (H2O) to reduce NOx emission [15]. In recent years, the molecular zeolite catalysts have been widely used to study the reduction of NO in NH3-SCR system [16] because of their unique microporous structure, rich acid sites, and strong thermal stability [17]. ZSM-5 molecular zeolite has a unique three-dimensional pore structure, large specific surface area, and strong NOx adsorption capacity. It shows excellent catalytic performance in NOx emission control of diesel engines and automobile exhaust [18]. However, the selective catalytic reduction activity of a single molecular zeolite is poor, which cannot meet the requirements of vehicle SCR technology.
At present, many experts and scholars unanimously seek better methods to improve SCR technology, especially in SCR catalyst, catalyst surface reaction mechanism, urea atomization, aftertreatment device design, and catalyst preparation. However, it is difficult to replace SCR catalyst in engine bench or vehicle test, the test conditions are poor and the extensive human and material resources invested. Therefore, the study of small catalyst samples in the SCR simulation reaction system can accurately control the reaction parameters of SCR catalytic reaction, such as temperature, space velocity, ammonia nitrogen ratio, and can conduct a more targeted and detailed study on the performance of the catalyst. Therefore, the most NH3-SCR uses zeolite molecular zeolite and transition metal groups (such as V [19], Cu [20], Fe [21] and Mn [22]) as catalysts for the reaction [23]. The vanadium-based catalyst is an early commercial deNOx catalyst with the advantages of high activity, high selectivity, and good sulfur resistance [24]. However, it has a narrow working window (350–450 °C), poor low-temperature activity, and poor high-temperature stability [25]. The Cu-exchanged zeolites are active both in the reduction of nitrogen oxides with NH3 in SCR system and in the direct decomposition of NO. In addition, Cu-exchanged zeolites SCR reactor includes a lot of parallel channels and mostly uses cordierite as substrate for its low cost, high thermal stability, and superior mechanical strength. Due to a relatively wide operating temperature range and the great low-temperature ignition characteristic, Cu zeolites are widely used in experimental research and application due to the high catalytic performance under normal vehicle exhaust temperatures. The heterogeneity of the reactive sites and numerous potential metal oxidation states that are prepared by incorporation of metal atoms into the zeolite framework result in complex zeolite catalysts. It can be found that Cu-zeolites active sites are various forms of copper oxides, Cu2+ and Cu+, or inherent Bronsted acid sites that present in the pores of the zeolites [26]. In 1986, Iwamoto et al. [27] found that the activity of Cu-ZSM-5 catalyst formed by exchanging Cu ions into ZSM-5 molecular sieve is much higher than that formed by exchanging Cu ions into other types of molecular sieves. In the following decades, researchers have carried out much research on catalytic decomposition by Cu-ZSM-5 [28].
Many experimental and numerical studies had been carried out on SCR coated with Cu-zeolites. For example, Na et al. [29] derived a nonlinear model based on the physical and chemical interpretation of the catalyst and some simplifications. Then, they discretized and linearized the nonlinear equations and analyzed the system dynamics of the catalyst. Through control, the maximum NOx conversion could be achieved with minimum NH3 leakage, and the model’s performance was verified by the SCR simulation reaction system. Zhong et al. [30] established a single channel NH3-SCR model of SCR system with Cu-ZSM-5 as a catalyst to study the mass and heat transfer processes between gas and solid phases. It was found that the deNOx efficiency was mainly influenced by intra-phase diffusion at 200–500 °C, while it was mainly influenced by inter-phase mass transfer at medium and high temperatures. Colombo et al. [31] experimentally studied the reaction law of Cu-based molecular zeolite NH3-SCR and analyzed the conversion of NOx in detail. Auvray et al. [32] established an axial spatial kinetic model of Cu-based molecular zeolite NH3-SCR reaction and analyzed the influence of NH3-SCR catalytic reaction. They found that pore diffusivity had little effect on it. In addition, Henry et al. [33] studied the axial spatial distribution of Cu-based molecular Auvray et al. [32] and established an axial spatial kinetic model of Cu-based molecular zeolite NH3-SCR reaction and analyzed the influence of NH3-SCR catalytic reaction. They analyzed the fast and standard SCR reaction conversion at each length position in the axial direction. Many accurate and effective experiment data can be obtained by simulating the engine test bench and evaluating catalyst performance by an SCR simulation reaction system. Compared with the engine bench test, it can eliminate the interference of the external environment and is more conducive to the systematic and scientific study of the reaction kinetic model of the catalyst.
Based on above, it is of great significance to study the influence of the Cu-ZSM-5 catalyst reaction kinetic model on the SCR system. Thus, a 3D- computational fluid dynamics model is established and employed to investigate the reaction kinetic model of Cu molecular sieve catalytic. In addition, the relative pressure drop, NH3 conversion, and NOx conversion efficiency in different channels were analyzed to provide some reference for improving, optimizing, and applying the NH3-SCR system. Thus, the work is very interesting.

2. Methods and Model Validation

2.1. SCR Reaction Kinetic Model

The SCR catalytic reactor has many parallel small-sized channels, a large channel length-width ratio, and a large number of channels. In the process of calculation, it is assumed that the flow field of the inlet gas of the SCR reactor is evenly distributed, and the heat transfer loss of the system to the environment is ignored. As shown in Figure 1, the phenomena of SCR mainly involves axial convection and diffusive transport in the gas phase, mass transfer through the gas-solid interface, and radial diffusion and catalytic reactions on the coating. When the NH3, NOx, and O2 in the channel enter the coating (interphase transfer), NH3 is adsorbed on the catalyst’s active sites and subsequently reacts with NOx and O2 for reduction. The remaining NH3, NOx, and O2 enter the coating (in-phase mass transfer) and continue to react with each other. At the same time, the generated H2O, N2, and other substances finally diffuse into the channels through the gaps. In the process of establishing the model, the mass transfer at the gas-solid interface adopts Cauchy boundary conditions, and the Hawthorn [34] equation calculates its mass transfer coefficient. At the same time, the model makes the following assumptions:
  • Ignore the diffusion between the whole channel, the hydrolysis of isocyanate and urea. NH3 directly enters the channel.
  • Assuming that the inlet boundary condition of the channel is uniform, the Eley-Rideal mechanism is used to calculate the surface catalytic reaction of the Cu molecular sieve.
  • The active catalyst is evenly coated on the coating surface, and the porous medium of the coating is evenly distributed.
The continuity equation, momentum equation, and enthalpy balance equation of gas-phase components in the channel can be expressed as follows [30]:
( ρ y i ) t + ( ρ v y i ) z   =   D i 2 ρ y i z 2
( ρ v ) t + ( ρ v 2 ) z   =   μ 2 v z 2 p z
c p ( ρ T g ) t + c p ( ρ v T g ) z   =   λ g 2 T g z 2
where ρ is the fluid density, kg/m3; y is the mass fraction of component, %; v is the volume, m3; t is time, s; z is the axial distance, m; Di is the diffusion coefficient of the species i in pores, m2/s; μ is the viscosity, Pa·s; cp is the specific heat capacity, J/(kg·K); T is the temperature, K; and λ is the thermal conductivity, W/(m·K).
The composition continuity equation, enthalpy balance equation, and NH3 coverage balance equation of gaseous substances in the coating are as follows [30]:
D w , i ρ w y w , i x 2 a c M i j γ i , j r j   =   ( ρ w y w , i ) t
a c j Δ H m , j r j ( T ) + λ w 2 T w x 2   =   c p ( ρ w T w ) t
Ω θ N H 3 t   =   j = 1 J γ N H 3 , j r j
where Dw,i is the effective diffusion coefficient of component i, m2/s; x is the mole fraction of component, %; yw,i is the mass fraction of component i, %; ac is the geometric surface area m2/m3; Mi is the molar mass of each i component, g/mol; rj is the reaction rate of reaction j, mol/(s·m3); ΔHm,j is the enthalpy of reaction j, J/mol; Ω is NH3 storage capacity, mol/m3; θNH3 is the NH3 surface site fraction; and γi,j is the stoichiometric coefficient of the reaction of component i at j.
The component continuity equation and enthalpy balance equation between gas and coating are as follows [30]:
D w , k y w , k x | x   =   ± d h 2   =   k m ( y b , k y w , k )
λ w T w x | x   =   ± d h 2   =   h T a c ( T g T w )
where km is the mass transfer coefficient, kg/(m2·s); hT is the heat transfer coefficient, kg/(s·m2); dh is the channel length, m; and k is the reaction constant, S−1.
The main reactions of the NH3-SCR catalytic reduction reaction are shown in Table 1.
In addition, the NH3 desorption activation energy parameter is modified by Temkin, and the following equation can calculate the NH3 adsorption and desorption reaction rates:
r a d s   =   k a d s c N H 3 ( θ max . N H 3 θ N H 3 )
r d e s   =   k d e s θ N H 3
where rdes and rads is the reaction rates of NH3 desorption and adsorption, mol/(s · m3); kdes and kads is the reaction rate constant of NH3 desorption and adsorption, s−1; cNH3 is the NH3 molar concentration, mol/m3.
The NO and NH3 reaction rates can be obtained by the following equation:
r N H 3   =   k N H 3 c o 2 θ N H 3
r N O   =   k N O ( C N O 1 / 2 c o 2 c N O / k e q )
where rNH3 and rNO is the reaction rates of NH3 and NO, mol/(s · m3); keq, kNH3 and kNO is the reaction rate constants of equilibrium, NH3 and NO, s−1; cNO and cO2 is the NO and O2 molar concentration, mol/m3.
The rapid and stand reaction rates can be obtained by the following equation:
r r a p   =   k r a p c N O c N O 2 θ N H 3 * [ 1 exp ( θ N H 3 / θ N H 3 * ) ]
r s t a   =   k s t a c N O θ N H 3 * [ 1 exp ( θ N H 3 / θ N H 3 * ) ]
where rrap and rsta is the rapid and stand reaction rates, mol/(s · m3); krap and ksta is the rapid and stand reaction rate constants, s−1; cNO2 is the NO2 molar concentration, mol/m3, θ*NH3 is the critical NH3 surface concentration value.

2.2. Pore Diffusion Model

Diffusion phenomena in pore media are widely present in external mass transfer and internal mass transfer processes in porous media. The diffusion of components in the diesel exhaust gas in SCR porous media is related to the micropores’ pore size, length, and shape [35]. The multi-component diffusion coefficient is obtained by the following equation:
D i m b u l k   =   1 x i j = 1 N j i x j D i j
The molecular diffusion coefficient of binary bulk can be obtained by the following equation [35]:
D i j b u l k   =   3.198 × 10 8 T 1.75 P ( V i 1 / 3 + V j 1 / 3 ) ( 1 M i + 1 M j ) 0.5
The Knudsen diffusion coefficient can be obtained by the following equation [36]:
D k   =   48.5 d p T M
The tortuosity square or tortuosity factor enters into the final form of the effective coefficient. In addition, the increase of superficial flux and the additional path length are considered. Based on the residence time of component i in the porous structure, a relation between the effective and pore gas diffusion coefficients is obtained by the following equation [37]:
D i e f f   =   ε D i , b u l k ( l e / l ) 2   =   ε τ 2 D i , b u l k
where τ is the tortuosity factor; l is the actual length of coating hole, m; le is the total length of coating, m; ε is the porosity of washcoat.
Based on the stationary and not-stationary diffusion, the Equation (18) is also confirmed in Ref. [38]. The porous structure is showed in Figure 2. The porous structure includes many cylindrical pores with the same cross-sectional area S. The length of each pore can be obtained by the following equation [39].
l e   =   d l e   =   d l cos θ
The above equation is reduce to le = l/cosθ. If there are n such pores per unit surface of the porous material, the porosity ε is obtained by the equation [30]:
ε   =   n S l e l
where S is cross-sectional area, m2.
Therefore, the total macroscopic molar diffusion flux in porous media in the sinusoidal capillary model can be calculated by the following equation [38]:
N   =   n S D b u l k c l e   =   ε D b u l k ( l e / l ) 2 c l   =   D e f f c l
where Dbulk is the bulk molecular diffusion coefficient; the Δc is the concentration difference of i component; the equation gives D e f f   =   ε D b u l k ( l e / l ) 2 , where τ   =   ( l e / l ) 2 is the tortuosity factor.
The tortuosity square can be employed to estimate the gas diffusion coefficients. A sudden change of any species concentration is put on the left-hand side in Figure 2, and the concentration on the right-hand side will start to increase after a certain time. The time can be obtained by the following equation [39]:
t   =   ( l e ) 2 / D b u l k l 2 / D e f f
According to the characteristics of the research object, the diffusion coefficients in the mesopores D i m e f f is a combination of the bulk and Knudsen diffusivities. Thus, the multi-component effective gas diffusion coefficient can be calculated by the following equation [38]:
D i m e f f   =   ε 2 ( 1 D i m b u l k + 1 D k ) 1
where Dimbulk is the diffusion coefficient of multi-component in the channel, m2/s; Dk is Knudsen diffusion coefficient, m2/s; Dimeff is the effective diffusion coefficient of multi-component, m2/s.
According to the conclusions and practical application of Ref. [40], the porosity value selects ε = τ−2/3 [41]. The calculated effective diffusion coefficients of reaction products O2, NO and NH3 are shown in Figure 3. It can be found that the model can accurately calculate the effective diffusion coefficient of each component.

2.3. Model Validation

In order to verify the accuracy of the simulation model, the Cu-ZSM-5 molecular sieves catalytic reactor was used to verify the model. The main parameters of the Cu-based molecular sieve catalytic reactor and catalyst are shown in Table 2. The AVL AMAi60 gas analyzer was used for the measurement. The different temperatures, flows, and pressures were measured by the corresponding sensors. The specific test bench is shown in Figure 4. The bench can generate The NH3 of 500 × 10−6, NO of 500 × 10−6, 2% H2O and 5% O2 (GHSV = 7200 h−1) can be generated by the experimental bench and the intake air can be controlled by the adjust valve. Then the gas is heated by a heater, and finally, a catalytic reaction occurs in the SCR catalyst. In order to prevent the cross-contamination of catalysts, all catalysts were first run with 5% O2 and N2 for 20 min at 500 °C to completely remove residual NOx and NH3 before the experiments were conducted. In addition, the pipes were insulated with glass wool to reduce heat loss.
The SCR reaction kinetic simulation model of Cu molecular zeolite catalyst was established in this paper and included mainly seven steps of the total reaction. The simulation results were optimized by adjusting the kinetic parameters such as pre-factor and activation energy in the model until the experimental results agreed with the simulation results. At the outlet, the experiment and simulation NO and NH3 concentrations in the standard SCR oxidation reaction are shown in Figure 5. It could be seen that the concentration errors of NO and NH3 at the outlet of experimental measurement and simulation calculation were within 5%, which indicated that the model could accurately predict the SCR reaction.

3. Results and Discussion

3.1. Axial Distribution of SCR Catalytic Reaction

NH3 with molar concentrations of 500 × 10−6, NO with molar concentrations of 500 × 10−6, 2% H2O, and 5% O2 were added to the SCR reactor. The NH3 concentration, NOx concentration, and NOx conversion in the axial direction of the catalyst under inlet conditions were simulated and studied in the steady-state standard SCR reaction, rapid SCR reaction, and NO2-SCR reaction [42,43].

3.1.1. Standard SCR Reaction

(1)
NH3 concentration
The variation of axial NH3 concentration under standard SCR reaction condition at a steady state is shown in Figure 6. It can be seen that the concentration of NH3 changes little at low temperatures. More specifically, with the variation of catalyst position, the effect is not obvious at 180 °C. It is due to the lower low-temperature activity of the Cu zeolite catalyst. When the temperature of the reactor exceeds 280 °C, the NH3 concentration gradually decreases with the increase of catalyst position. This is due to the higher low-temperature activity of the Cu zeolite catalyst at high temperatures. Thus, the concentration of NH3 decreases with the increase of catalyst position. In addition, Komatsu et al. [44] suggested that Cu particles increased the concentration of NO2, which reacted with NH3 and produced N2 and H2O, thus reducing the NH3 concentration. Thus, the higher the temperature, the more NH3 concentration decreases. Auvary et al. [32] and Gong et al. [45] have similar conclusions.
(2)
NOx concentration
Figure 7 shows the variation of catalyst axial NOx concentration under standard SCR reaction condition at a steady state. It can be seen that the NOx concentration changes little at low temperatures. With the increase of temperature, the NOx conversion increases due to the increasing gradually activity of the Cu zeolite catalyst. Between 380 °C and 480 °C, NOx on the Cu zeolite catalyst surface can be almost completely converted, resulting in the decrease of NOx concentration. More specifically, at catalyst position of 1.0 cm, the conversion efficiency has reached more than 85%. When the catalyst position is 1.9 cm, 95% of the NOx has been converted. Henry et al. [33] had a similar conclusion. Therefore, the length of the catalyst can be reasonably selected.
(3)
NOx conversion
The variation of axial NOx conversion under standard SCR reaction conditions at a steady state are shown in Figure 8. It can be seen that the concentration of NOx changes little at low temperatures. Only when the temperature in the reactor exceeds 280 °C, the activity of the catalyst increases significantly. More specifically, when the temperature reaches 380 °C, the conversion rate has reached more than 95% at 1.2 cm. Metkar et al. [46] and Xue et al. [47] also proved that the higher the temperature, the higher the NOx conversion efficiency. In addition, it can be found that the SCR reaction is mainly concentrated in the first half part of the reactor, and the influence of the rear section is small. Therefore, the length of the reactor can be reasonably selected to optimize its structure.

3.1.2. Fast SCR Reaction

(1)
NH3 concentration
The variation of NH3 concentration in the axial direction of the catalyst under fast SCR reaction conditions at a steady state is shown in Figure 9. Compared with standard SCR reaction (Seen Figure 6), the NH3 concentration of fast SCR reaction is also significantly transformed at low temperatures. Moreover, the NH3 concentration decreases faster in the first half of the catalyst. In addition, the concentration of NH3 is lower than that of NOx. This is due to the reduction of NH3 concentration due to the existence of an NH3 oxidation reaction. Metkar et al. [48] and Lu et al. [49] have similar experimental conclusions.
(2)
NOx concentration
Figure 10 shows the catalyst axial NOx concentration variation under the fast SCR reaction condition in a steady state. Compared with the standard SCR, NOx in the first half of the reactor is also significantly converted in the fast SCR reaction, and the speed of the fast SCR reaction is faster. This is because the NOx conversion activity of the zeolite catalyst gradually increases with the increase of temperature under the fast reaction. Activity has been greatly improved. Meanwhile, the concentration of NOx in the reactor is higher than that of NH3 due to an NH3 oxidation reaction (Figure 9).
(3)
NOx conversion
Figure 11 shows the catalyst axial NOx conversion variation under the fast SCR reaction condition in a steady state. It can be seen that the NOx conversion rate of the fast SCR reaction increases gradually at all temperatures. It is because the presence of NO2 in the reaction improves the NOx conversion efficiency and the reaction is much faster than the standard SCR reaction. In addition, with the increase of catalyst position, the conversion increases gradually. This shows that the activity of Cu-based catalysts is enhanced under fast reaction. At the same temperature, the position dependence of the rapid SCR reaction on the catalyst decreases (after 0.9 cm, the effect is small), which may be due to the higher total conversion caused by the mixed reaction control.

3.1.3. NO2-SCR Reaction

(1)
NOx concentration
The variation of catalyst axial NO2 concentration under NO2-SCR reaction condition in a steady state are shown in Figure 12. It can be seen that more NO2 will be produced at 180–280 °C, and the concentration of NO2 increases with the increase of reactor location. However, there is less NO2 in the high-temperature stage. This is because the oxidation reaction of NO is reversible. At low temperature, the reaction rate is controlled by reaction kinetics, while at high temperature, the reaction is controlled by reaction equilibrium, and NO2 decomposes to produce NO. For example, Shi et al. [50] reached a similar conclusion. In addition, it can be found that the concentration of NO2 remains unchanged when the reactor position is 0.6 cm at high temperatures (380–480 °C).
(2)
NOx conversion
The variation of catalyst axial NOx conversion under NO2-SCR reaction condition in a steady state are shown in Figure 13. It can be seen that the presence of NO2 improves the overall conversion of NOx. However, NO2-SCR reaction is faster than standard SCR but slower than the fast SCR. In addition, NO2-SCR reaction is mainly concentrated in the first half of the catalytic reactor, and there is only a slight increase in the second half. Moreover, at 280–480 °C, the NOx conversion rate is relatively high. This shows that the catalytic activity increases gradually with the increase of temperature. NOx reacts from 280 °C, and the conversion rate increases gradually with the increase of the position of the catalyst. Finally, it tends to be stable. In addition, it can be seen from Figure 13 that the conversion efficiency of NOx is the best at the position of 1.2 cm.

3.2. Reaction Conditions of Different SCR Channel Cross-Sections

In order to further improve the heat and mass transfer of the SCR channel, three different structures of SCR channels were designed and then calculated and simulated to optimize the performance of each SCR channel. The first structure is a square channel, the second structure is a rectangular channel, and the third structure is a hexagonal channel. The wall thickness C of the three channels is 0.7 mm, and the length A and width B are shown in Figure 14. Under the same boundary condition, the effects of different length-width ratio structures on SCR channel performance are studied using the verified simulation model in term of pressure drop, NH3 conversion, and NOx conversion are discussed.
Based on the calculation results of square (A = 4.0 mm), rectangular channels with different width B (3.0–3.6 mm) and different length-width ratios A/B (1–10) are selected in the paper. The established simulation model is used for simulation calculation, and the NO conversion, NH3 conversion, and pressure drop results of the relative square channel (A = 4.0 mm) with the same wall thickness and different rectangular channels are obtained, as shown in Figure 15a–c.
It can be seen that with the increase of length-width ratio, the relative pressure drop, relative NO conversion, and relative NH3 conversion all show a decreasing trend. When the length-width ratio is greater than 5, the length-width ratio has little effect on the relative pressure drop, relative NO conversion, and NH3 conversion. This also shows that when selecting rectangular or square channels, the interval with a length-width ratio of 1–5 is a range worthy of consideration. For example, compared with square (A = 4.0), when the length-width ratio is 2, the relative pressure drop, relative NH3 conversion, and relative NO conversion ratio of the rectangle (A = 3.2) are reduced by 6.8%, 0.15%, and 0.12%, respectively. Meanwhile, with the increase of width A, the relative pressure drop, relative NO conversion, and relative NH3 conversion decreased.
A significant disadvantage of square and rectangular channels is that the adjacent edges intersect at a right angle. It is unfavorable for gas flow. Thus, the gas pressure drop is increased, and the reduction efficiency of nitrogen oxides is reduced in the channel. In addition, sharp right-angle edges are easy to be block, resulting in the failure of SCR porous media channel. In order to overcome this deficiency, this paper changes the rectangular channel with 90° edge to the hexagonal channel with 120° edge. The chemical reaction of gas will be improved due the improvement of gas flow. The velocity cross-section distribution of gas passing through the rectangular and hexagonal channels under the same boundary conditions is shown in Figure 16. It can be seen that the gas flow field distribution in the hexagonal channel is more uniform than that in the rectangular channel. At the corner of the rectangular channel, the flow rate of a large part of the reaction gas is very low or even stagnant. Therefore, more dead zones of gas reaction with catalyst are formed at the corner of the rectangular channel. Among them, the area where the cross-section velocity of the rectangular channel is less than 5 m/s accounts for 38%. Compared with rectangular channel, the area where the cross-section velocity of the hexagonal channel is less than 5 m/s accounts for 33%. Thus, the hexagonal channel can effectively increase the diffusion rate of gas reactants to the catalyst wall to improve the catalytic conversion efficiency.
The relative NO conversion, relative NH3 conversion, and relative pressure drop for the rectangular and hexagonal channels relative to the square channel are shown in Figure 17a–c. It can be found that the hexagonal channel further enhances the heat and mass transfer between the gas and the catalyst wall and reduces the pressure drop of the gas in the channel when the length-width ratio and wall thickness are constant. For example, compared with the square channel (A = 4.0), the relative pressure drop of the designed hexagonal channel is reduced by 9.6%, and the relative NO conversion and relative NH3 conversion are increased by 0.51% and 0.48%. Therefore, a reasonable selection of structural parameters of SCR catalysts can effectively optimize the SCR catalyst system.

4. Conclusions

With the increasingly stringent emission regulations due to energy crisis [51,52,53,54,55,56,57] and environmental problems [58,59,60,61,62,63,64,65,66,67,68], how to effectively reduce NOx emission from a diesel engine has become a concern [69]. In order to reduce the NOx emission, the kinetic model of NH3-SCR reaction over Cu-ZSM-5 catalyst was established in this paper. Firstly, the NH3 concentration, NOx concentration, and NOx conversion rate variation laws along the axis were analyzed. Secondly, the relative pressure drops, relative NH3 conversions, and relative NOx conversion efficiencies of different SCR channels were studied. Finally, the main conclusions were as follows:
(1)
The improved model can better predict the catalytic reaction of the diesel SCR system.
(2)
The Cu-based zeolite catalyst is of poor low-temperature characteristics and good high-temperature activity.
(3)
NH3-SCR catalytic reaction is mainly concentrated in the first half of the reactor, and the influence of the second half is little.
(4)
The hexagonal channel can effectively increase the diffusion rate of gas reactants to the catalyst wall, reduce the pressure drop and improve the catalytic conversion efficiencies of NH3 and NOx.

Author Contributions

Conceptualization, S.Y. and J.Z.; software, S.Y.; formal analysis, S.Y. and J.Z.; investigation, S.Y. and J.Z.; resources, J.Z.; writing—original draft preparation, S.Y. and J.Z.; writing—review and editing, S.Y. and J.Z.; supervision, J.Z.; funding acquisition, J.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Natural Science Foundation of Jiangsu Province under the research grant BK20191205 in China; the work is supported by the Natural Science Foundation of Guangxi under the research grants of 2018GXNSFAA281267 and 2018 GXNSFAA 294072.

Data Availability Statement

All data used to support the findings of this study are included within the article.

Conflicts of Interest

The authors declare that they have no conflict of interests regarding the publication of this paper.

References

  1. Zhang, Z.; E, J.; Chen, J.; Zhu, H.; Zhao, X.; Han, D.; Zuo, W.; Peng, Q.; Gong, J.; Yin, Z. Effects of low-level water addition on spray, combustion and emission characteristics of a medium speed diesel engine fueled with biodiesel fuel. Fuel 2018, 239, 245–262. [Google Scholar] [CrossRef]
  2. Emiroğlu, A.O. Effect of fuel injection pressure on the characteristics of single cylinder diesel engine powered by butanol-diesel blend. Fuel 2019, 256, 115928. [Google Scholar] [CrossRef]
  3. Zhang, Z.; E, J.; Chen, J.; Zhao, X.; Zhang, B.; Deng, Y.; Peng, Q.; Yin, Z. Effects of boiling heat transfer on the performance enhancement of a medium speed diesel engine fueled with diesel and rapeseed methyl ester. Appl. Therm. Eng. 2020, 169, 114984. [Google Scholar] [CrossRef]
  4. Zhang, Z.; Ye, J.; Tan, D.; Feng, Z.; Luo, J.; Tan, Y.; Huang, Y. The effects of Fe2O3 based DOC and SCR catalyst on the combustion and emission characteristics of a diesel engine fueled with biodiesel. Fuel 2021, 290, 120039. [Google Scholar] [CrossRef]
  5. E, J.; Zhao, X.; Liu, G.; Zhang, B.; Zuo, Q.; Wei, K.; Li, H.; Han, D.; Gong, J. Effects analysis on optimal microwave energy consumption in the heating process of composite regeneration for the diesel particulate filter. Appl. Energy 2019, 254, 113736. [Google Scholar] [CrossRef]
  6. Di Benedetto, A. Modeling and simulation of soot combustion dynamics in a catalytic diesel particulate filter. Chem. Eng. Sci. 2015, 137, 69–78. [Google Scholar]
  7. Li, W.; Ji, J.; Huang, L.; Guo, Z. Global dynamics of a controlled discontinuous diffusive SIR epidemic system. Appl. Math. Lett. 2021, 121, 107420. [Google Scholar] [CrossRef]
  8. Rogóż, R.; Kapusta, J.; Bachanek, J.; Vankan, J.; Teodorczyk, A. Improved urea-water solution spray model for simulations of selective catalytic reduction systems. Renew. Sustain. Energy Rev. 2019, 120, 109616. [Google Scholar] [CrossRef]
  9. E, J.; Zhao, X.; Qiu, L.; Wei, K.; Zhang, Z.; Deng, Y.; Han, D.; Liu, G. Experimental investigation on performance and economy characteristics of a diesel engine with variable nozzle turbocharger and its application in urban bus. Energy Convers. Manag. 2019, 193, 149–161. [Google Scholar] [CrossRef]
  10. Zhang, Z.; E, J.; Deng, Y.; Pham, M.; Zuo, W.; Peng, Q.; Yin, Z. Effects of fatty acid methyl esters proportion on combustion and emission characteristics of a biodiesel fueled marine diesel engine. Energy Convers. Manag. 2018, 159, 244–253. [Google Scholar] [CrossRef]
  11. Tan, D.; Chen, Z.; Li, J.; Luo, J.; Yang, D.; Cui, S.; Zhang, Z. Effects of Swirl and Boiling Heat Transfer on the Performance Enhancement and Emission Reduction for a Medium Diesel Engine Fueled with Biodiesel. Processes 2021, 9, 568. [Google Scholar] [CrossRef]
  12. E, J.; Liu, G.; Zhang, Z.; Han, D.; Chen, J.; Gong, J.; Yin, Z. Effect analysis on cold starting performance of a diesel engine fueled with biodiesel fuel based on an improved thermodynamic model. Appl. Energy 2019, 243, 321–335. [Google Scholar] [CrossRef]
  13. Guan, B.; Zhan, R.; Lin, H.; Huang, Z. Review of state of the art technologies of selective catalytic reduction of NOx from diesel engine exhaust. Appl. Therm. Eng. 2014, 66, 395–414. [Google Scholar] [CrossRef]
  14. Cai, T.; Zhao, D. Effects of fuel composition and wall thermal conductivity on thermal and NOx emission performances of an ammonia/hydrogen-oxygen micro-power system. Fuel Process. Technol. 2020, 209, 106527. [Google Scholar] [CrossRef]
  15. Zhang, Z.; Tian, J.; Li, J.; Ji, H.; Tan, D.; Luo, J.; Jiang, Y.; Yang, D.; Cui, S. Effects of Different Mixture Ratios of Methanol-Diesel on the Performance Enhancement and Emission Reduction for a Diesel Engine. Processes 2021, 9, 1366. [Google Scholar] [CrossRef]
  16. Zhang, Z.; Li, J.; Tian, J.; Xie, G.; Tan, D.; Qin, B.; Huang, Y.; Cui, S. Effects of Different Diesel-Ethanol Dual Fuel Ratio on Performance and Emission Characteristics of Diesel Engine. Processes 2021, 9, 1135. [Google Scholar] [CrossRef]
  17. Wang, X.; Zhang, R.; Wang, H.; Wei, Y. Strategy on Effective Synthesis of SSZ-13 Zeolite Aiming at Outstanding Performances for NH3-SCR Process. Catal. Surv. Asia 2020, 24, 143–155. [Google Scholar] [CrossRef]
  18. Xue, H.; Guo, X.; Meng, T.; Guo, Q.; Mao, D.; Wang, S. Cu-ZSM-5 Catalyst Impregnated with Mn–Co Oxide for the Selected Catalytic Reduction of NO: Physicochemical Property–Catalytic Activity Relationship and In Situ DRIFTS Study for the Reaction Mechanism. ACS Catal. 2021, 11, 7702–7718. [Google Scholar] [CrossRef]
  19. Naicker, T.; Friedrich, H.B. Hexane activation over vanadium modified zeolite ZSM-5. J. Porous Mater. 2012, 20, 763–775. [Google Scholar] [CrossRef]
  20. Zhong, C.; Tan, J.; Zuo, H.; Wu, X.; Wang, S.; Liu, J. Synergy effects analysis on CDPF regeneration performance enhancement and NOx concentration reduction of NH3–SCR over Cu–ZSM–5. Energy 2021, 230, 120814. [Google Scholar] [CrossRef]
  21. E, J.; Zhang, Z.; Chen, J.; Pham, M.; Zhao, X.; Peng, Q.; Zhang, B.; Yin, Z. Performance and emission evaluation of a marine diesel engine fueled by water biodiesel-diesel emulsion blends with a fuel additive of a cerium oxide nanoparticle. Energy Convers. Manag. 2018, 169, 194–205. [Google Scholar] [CrossRef]
  22. Cheng, M.; Jiang, B.; Yao, S.; Han, J.; Zhao, S.; Tang, X.; Zhang, J.; Wang, T. The Mechanism of NH3-SCR Reaction for NOx Removal from Diesel Engine Exhaust and Hydrothermal Stability of Cu-Mn/Zeolite Catalysts. J. Phys. Chem. C 2017, 122, 1–35. [Google Scholar]
  23. Fu, G.; Chen, J.; Liang, Y.; Li, R.; Yang, X.; Jiang, J. Cu-IM-5 as the Catalyst for Selective Catalytic Reduction of NOx with NH3: Role of Cu Species and Reaction Mechanism. Catalysts 2021, 11, 221. [Google Scholar] [CrossRef]
  24. Rajasekar, V.; Ashok, B. Critical interpretative review on current outlook and prospects of selective catalytic reduction system for De-NOx strategy in compression ignition engine. Fuel 2020, 276, 117996. [Google Scholar]
  25. Grossale, A.; Nova, I.; Tronconi, E. Study of a Fe–zeolite-based system as NH3-SCR catalyst for diesel exhaust aftertreatment. Catal. Today 2008, 136, 18–27. [Google Scholar] [CrossRef]
  26. Chapman, D.M. Behavior of titania-supported vanadia and tungsta SCR catalysts at high temperatures in reactant streams: Tungsten and vanadium oxide and hydroxide vapor pressure reduction by surficial stabilization. Appl. Catal. A: Gen. 2011, 392, 143–150. [Google Scholar] [CrossRef]
  27. Iwamoto, M.; Furukawa, H.; Mine, Y.; Uemura, F.; Mikuriya, S.-I.; Kagawa, S. Copper(II) ion-exchanged ZSM-5 zeolites as highly active catalysts for direct and continuous decomposition of nitrogen monoxide. J. Chem. Soc. Chem. Commun. 1986, 1272–1273. [Google Scholar] [CrossRef]
  28. Zhou, D.; Yang, W.; An, H.; Li, J.; Shu, C. A numerical study on RCCI engine fueled by biodiesel/methanol. Energy Convers. Manag. 2014, 89, 798–807. [Google Scholar] [CrossRef]
  29. Na, H.; Reed, D.C.; Annaswamy, A.M.; Laing, P.M.; Kolmanovsky, I. Control-Oriented Reduced-Order Models for Urea Selective Catalytic Reduction Systems Using a Physics-Based Approach. In Proceedings of the SAE 2011 World Congress & Exhibition, Warrendale, PA, USA, 12–14 April 2011. [Google Scholar] [CrossRef]
  30. Zhong, C.; Gong, J.; Tan, L.; Liu, W.; Liu, G.; Zhang, Z. Modeling intraphase and interphase mass transfer limitations for NH3–SCR over Cu–ZSM–5. Chem. Eng. Sci. 2019, 207, 479–489. [Google Scholar] [CrossRef]
  31. Colombo, M.; Nova, I.; Tronconi, E. Detailed kinetic modeling of the NH3–NO/NO2 SCR reactions over a commercial Cu-zeolite catalyst for Diesel exhausts after treatment. Catal. Today 2012, 197, 243–255. [Google Scholar] [CrossRef]
  32. Auvray, X.; Partridge, W.; Choi, J.-S.; Pihl, J.; Yezerets, A.; Kamasamudram, K.; Currier, N.; Olsson, L. Local ammonia storage and ammonia inhibition in a monolithic copper-beta zeolite SCR catalyst. Appl. Catal. B-Environ. 2012, 126, 144–152. [Google Scholar] [CrossRef]
  33. Henry, C.; Kamasamudram, K.; Currier, N.W.; Yezerets, A.; Castagnola, M.; Chen, H.-Y. Axially Resolved Performance of Cu-Zeolite SCR Catalysts. SAE Int. J. Engines 2012, 5, 975–984. [Google Scholar] [CrossRef]
  34. Hawthorn, R.D. Afterburner catalysts effects of heat and mass transfer between gas and catalyst surface. Recent Adv. Air Pollut. 1974, 70, 428–438. [Google Scholar]
  35. Kong, W.; Zhu, H.; Fei, Z.; Lin, Z. A modified dusty gas model in the form of a Fick’s model for the prediction of multicomponent mass transport in a solid oxide fuel cell anode. J. Power Sources 2012, 206, 171–178. [Google Scholar] [CrossRef]
  36. Hayes, R.; Kolaczkowskib, S.; Li, P.; Awdry, S. Evaluating the effective diffusivity of methane in the washcoat of a honeycomb monolith. Appl. Catal. B Environ. 2000, 25, 93–104. [Google Scholar] [CrossRef]
  37. Epstein, N. On tortuosity and the tortuosity factor in flow and diffusion through porous media. Chem. Eng. Sci. 1989, 44, 777–779. [Google Scholar] [CrossRef]
  38. Yuan, J.; Sundén, B. On mechanisms and models of multi-component gas diffusion in porous structures of fuel cell electrodes. Int. J. Heat Mass Transf. 2014, 69, 358–374. [Google Scholar] [CrossRef]
  39. Gommes, C.J.; Bons, A.J.; Blacher, S.; Dunsmuir, J.H.; Tsou, A.H. Practical methods for measuring the tortuosity of porous materials from binary or gray-tone tomographic reconstructions. Am. Inst. Chem. Eng. 2009, 55, 2000–2012. [Google Scholar] [CrossRef]
  40. Zhang, F.; Hayes, R.; Kolaczkowski, S. A New Technique to Measure the Effective Diffusivity in a Catalytic Monolith Washcoat. Chem. Eng. Res. Des. 2004, 82, 481–489. [Google Scholar] [CrossRef]
  41. Hu, X.; Yang, M.; Fan, D.; Qi, G.; Wang, J.; Wang, J.; Yu, T.; Li, W.; Shen, M. The role of pore diffusion in determining NH3 SCR active sites over Cu/SAPO-34 catalysts. J. Catal. 2016, 341, 55–61. [Google Scholar] [CrossRef]
  42. Santos, A.; Bahamonde, A.; Avila, P.; Garcia-Ochoa, F. Measurement of the effective diffusivity for a vanadia-tungsta-titania/sepiolite catalyst for SCR of NOx. Appl. Catal. B Environ. 1996, 8, 299–314. [Google Scholar] [CrossRef]
  43. Popovičová, J.; Brusseau, M.L. Dispersion and transport of gas-phase contaminants in dry porous media: Effect of heterogeneity and gas velocity. J. Contam. Hydrol. 1997, 28, 157–169. [Google Scholar] [CrossRef]
  44. Komatsu, T.; Nunokawa, M.; Moon, I.; Takahara, T.; Namba, S.; Yashima, T. Kinetic Studies of Reduction of Nitric Oxide with Ammonia on Cu2+-Exchanged Zeolites. J. Catal. 1994, 148, 427–437. [Google Scholar] [CrossRef]
  45. Gong, J.; Narayanaswamy, K.; Rutland, C.J. Heterogeneous Ammonia Storage Model for NH3-SCR Modeling. Ind. Eng. Chem. Res. 2016, 55, 5874–5884. [Google Scholar] [CrossRef]
  46. Metkar, P.S.; Balakotaiah, V.; Harold, M.P. Experimental study of mass transfer limitations in Fe- and Cu-zeolite-based NH3-SCR monolithic catalysts. Chem. Eng. Sci. 2011, 66, 5192–5203. [Google Scholar] [CrossRef]
  47. Xue, H.; Meng, T.; Liu, F.; Guo, X.; Wang, S.; Mao, D. Enhanced resistance to calcium poisoning on Zr-modified Cu/ZSM-5 catalysts for the selective catalytic reduction of NO with NH3. RSC Adv. 2019, 9, 38477–38485. [Google Scholar] [CrossRef] [Green Version]
  48. Metkar, P.S.; Harold, M.P.; Balakotaiah, V. Experimental and kinetic modeling study of NH3-SCR of NOx on Fe-ZSM-5, Cu-chabazite and combined Fe- and Cu-zeolite monolithic catalysts. Chem. Eng. Sci. 2013, 87, 51–66. [Google Scholar] [CrossRef]
  49. Lu, Y.; Jiang, Z.; Geng, N.; Jiang, S.; Xie, X. Appointment window scheduling with wait-dependent abandonment for elective inpatient admission. Int. J. Prod. Res. 2021, 1–17. [Google Scholar] [CrossRef]
  50. Shi, L.; Si, Z.; Wu, X.; Weng, D.; Wu, Z. Effect of water vapor on NH3–NO/NO2 SCR performance of fresh and aged MnOx–NbOx–CeO2 catalysts. J. Environ. Sci. 2015, 31, 240–247. [Google Scholar]
  51. Zuo, H.; Tan, J.; Wei, K.; Huang, Z.; Zhong, D.; Xie, F. Effects of different poses and wind speeds on wind-induced vibration characteristics of a dish solar concentrator system. Renew. Energy 2021, 168, 1308–1326. [Google Scholar] [CrossRef]
  52. Zuo, H.; Liu, G.; Jiaqiang, E.; Zuo, W.; Wei, K.; Hu, W.; Tan, J.; Zhong, D. Catastrophic analysis on the stability of a large dish solar thermal power generation system with wind-induced vibration. Sol. Energy 2019, 183, 40–49. [Google Scholar] [CrossRef]
  53. Hu, L.; Hu, X.; Che, Y.; Feng, F.; Lin, X.; Zhang, Z. Reliable state of charge estimation of battery packs using fuzzy adaptive federated filtering. Appl. Energy 2020, 262, 114569. [Google Scholar] [CrossRef]
  54. Zhang, F.; Liao, G.; E, J.; Chen, J.; Leng, E. Comparative study on the thermodynamic and economic performance of novel absorption power cycles driven by the waste heat from a supercritical CO2 cycle. Energy Convers. Manag. 2021, 228, 113671. [Google Scholar] [CrossRef]
  55. Zuo, H.; Zhang, B.; Huang, Z.; Wei, K.; Tan, J. Effect analysis on SOC values of the power lithium manganate battery during discharging process and its intelligent estimation. Energy 2022, 238, 121854. [Google Scholar] [CrossRef]
  56. E, J.; Zhang, B.; Zeng, Y.; Wen, M.; Huang, Z.; Wei, K.; Chen, J.; Zhu, H.; Deng, Y. Effects analysis on active equalization control of lithium-ion batteries based on intelligent estimation of the state-of-charge. Energy 2022, 238, 121822. [Google Scholar] [CrossRef]
  57. E, J.; Luo, B.; Han, D.; Chen, J.; Liao, G.; Zhang, F.; Ding, J. A comprehensive review on performance improvement of micro energy mechanical system: Heat transfer, micro combustion and energy conversion. Energy 2022, 240, 122509. [Google Scholar] [CrossRef]
  58. Ma, Y.; Liu, C.; E, J.; Mao, X.; Yu, Z. Research on modeling and parameter sensitivity of flow and heat transfer process in typical rectangular microchannels: From a data-driven perspective. Int. J. Therm. Sci. 2022, 172, 107356. [Google Scholar] [CrossRef]
  59. Yi, F.; E, J.; Zhang, B.; Zuo, H.; Wei, K.; Chen, J.; Zhu, H.; Zhu, H.; Deng, Y. Effects of coupling phase change material and liquid cooling on heat dissipation characteristics of Lithium-ion battery thermal management system. Renewable Energy 2022, 181, 472–489. [Google Scholar] [CrossRef]
  60. Cai, T.; Sun, Y.; Zhao, D. Enhancing heat transfer performance analyses of a hydrogen-fueled meso-combustor with staggered bluff-bodies. Fuel Process. Technol. 2021, 218, 106867. [Google Scholar] [CrossRef]
  61. Cai, T.; Zhao, D. Mitigating NOx emissions from an ammonia-fueled micro-power system with a perforated plate implemented. J. Hazard. Mater. 2021, 401, 123848. [Google Scholar] [CrossRef] [PubMed]
  62. Cai, T.; Zhao, D.; Wang, B.; Li, J.W.; Guan, Y.H. NOx emission and thermal performances studies on premixed ammonia-oxygen combustion in a CO2-free micro-planar combustor. Fuel 2020, 280, 118554. [Google Scholar] [CrossRef]
  63. E, J.; Pham, M.; Zhao, D.; Deng, Y.; Le, D.; Zuo, W.; Zhu, H.; Liu, T.; Peng, Q.; Zhang, Z. Effect of different technologies on combustion and emissions of the diesel engine fueled with biodiesel: A review. Renew. Sustain. Energy Rev. 2017, 80, 620–647. [Google Scholar] [CrossRef]
  64. Zhang, B.; Zuo, H.; Huang, Z.; Tan, J.; Zuo, Q. Endpoint forecast of different diesel-biodiesel soot filtration process in diesel particulate filters considering ash deposition. Fuel 2020, 272, 117678. [Google Scholar] [CrossRef]
  65. E, J.; Zhao, M.; Zuo, Q.; Zhang, B.; Zhang, Z.; Peng, Q.; Han, D.; Zhao, X.; Deng, Y. Effects analysis on diesel soot continuous regeneration performance of a rotary microwave-assisted regeneration diesel particulate filter. Fuel 2020, 260, 116353. [Google Scholar] [CrossRef]
  66. Feng, C.; Deng, Y.; Chen, L.; Han, W.; E, J.; Wei, K.; Han, D.; Zhang, B. Hydrocarbon emission control of a hydrocarbon adsorber and converter under cold start of the gasoline engine. Energy 2022, 239, 122138. [Google Scholar] [CrossRef]
  67. Chen, L.; Deng, Y.; Feng, C.; Han, W.; E, J.; Wang, C.; Han, D.; Zhang, B. Effects of zeolite molecular sieve on the hydrocarbon adsorbent performance of gasoline engine of during cold start. Fuel 2022, 310, 122427. [Google Scholar] [CrossRef]
  68. E, J.; Luo, J.; Han, D.; Tan, Y.; Feng, C.; Deng, Y. Effects of different catalysts on light-off temperature of volatile organic components in the rotary diesel particulate filter during the regeneration. Fuel 2022, 310, 122451. [Google Scholar] [CrossRef]
  69. Cai, T.; Zhao, D.; Sun, Y.; Ni, S.; Li, W. Evaluation of NOx Emissions Characteristics in a CO2-Free Micro-Power System by Implementing a Perforated Plate. Renew. Sustain. Energy Rev. 2021, 145, 111150. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of single-channel in selective catalytic reduction (SCR).
Figure 1. Schematic diagram of single-channel in selective catalytic reduction (SCR).
Processes 09 01966 g001
Figure 2. Simplified coating model.
Figure 2. Simplified coating model.
Processes 09 01966 g002
Figure 3. Effective diffusion coefficients of O2, NO, and NH3, in the coating.
Figure 3. Effective diffusion coefficients of O2, NO, and NH3, in the coating.
Processes 09 01966 g003
Figure 4. Schematic diagram of diesel engine test device.
Figure 4. Schematic diagram of diesel engine test device.
Processes 09 01966 g004
Figure 5. Comparison of NO and NH3 concentrations calculated by experiment and simulation at the reaction outlet.
Figure 5. Comparison of NO and NH3 concentrations calculated by experiment and simulation at the reaction outlet.
Processes 09 01966 g005
Figure 6. Variation of axial NH3 concentration under standard SCR reaction.
Figure 6. Variation of axial NH3 concentration under standard SCR reaction.
Processes 09 01966 g006
Figure 7. Variation of axial NOx concentration under standard SCR reaction.
Figure 7. Variation of axial NOx concentration under standard SCR reaction.
Processes 09 01966 g007
Figure 8. Variation of axial NOx conversion under standard SCR reaction.
Figure 8. Variation of axial NOx conversion under standard SCR reaction.
Processes 09 01966 g008
Figure 9. Variation of axial NH3 concentration under fast SCR reaction.
Figure 9. Variation of axial NH3 concentration under fast SCR reaction.
Processes 09 01966 g009
Figure 10. Variation of axial NOx concentration under fast SCR reaction.
Figure 10. Variation of axial NOx concentration under fast SCR reaction.
Processes 09 01966 g010
Figure 11. Variation of axial NOx conversion under fast SCR reaction.
Figure 11. Variation of axial NOx conversion under fast SCR reaction.
Processes 09 01966 g011
Figure 12. Variation of axial NOx concentration under NO2-SCR reaction condition.
Figure 12. Variation of axial NOx concentration under NO2-SCR reaction condition.
Processes 09 01966 g012
Figure 13. Variation of axial NOx conversion under NO2-SCR reaction condition.
Figure 13. Variation of axial NOx conversion under NO2-SCR reaction condition.
Processes 09 01966 g013
Figure 14. Sectional area dimension of SCR channel (length-width ratio = A/B, C is wall thickness).
Figure 14. Sectional area dimension of SCR channel (length-width ratio = A/B, C is wall thickness).
Processes 09 01966 g014
Figure 15. Relative NO conversion, relative NH3 conversion, and relative pressure drop changes relative to square channel and rectangular channel, (a) Relative NO conversion, (b) Relative NH3 conversion, (c) Relative pressure drop.
Figure 15. Relative NO conversion, relative NH3 conversion, and relative pressure drop changes relative to square channel and rectangular channel, (a) Relative NO conversion, (b) Relative NH3 conversion, (c) Relative pressure drop.
Processes 09 01966 g015
Figure 16. Velocity distribution of rectangular and hexagonal channels with the same wall thickness.
Figure 16. Velocity distribution of rectangular and hexagonal channels with the same wall thickness.
Processes 09 01966 g016
Figure 17. Relative NO conversion, NH3 conversion, and pressure drop for rectangular and hexagonal channels relative to square channels, (a) Relative NO conversion, (b) Relative NH3 conversion, (c) Relative pressure drop.
Figure 17. Relative NO conversion, NH3 conversion, and pressure drop for rectangular and hexagonal channels relative to square channels, (a) Relative NO conversion, (b) Relative NH3 conversion, (c) Relative pressure drop.
Processes 09 01966 g017aProcesses 09 01966 g017b
Table 1. Catalytic reaction of Cu molecular sieve.
Table 1. Catalytic reaction of Cu molecular sieve.
NO.Reactions
R1 N H 3 + S 1 N H 3 S 1
R2 N H 3 S 1 N H 3 + S 1
R3 2 N O O 2 2 N O 2
R4 2 N H 3 S 1 + N O 2 + N O 2 N 2 + 3 H 2 O + 2 S 1
R5 4 N H 3 S 1 + 4 N O + O 2 4 N 2 + 6 H 2 O + 4 S 1
R6 4 N H 3 S 1 + 3 O 2 2 N 2 + 6 H 2 O + 4 S 1
Table 2. Reactor and catalyst parameters.
Table 2. Reactor and catalyst parameters.
Parameter/UnitValue
Cell density/cell/in−2400
Coating thickness/mm0.025
Carrier length/mm20
Temperature/K373–873
Cu loading/wt%2
Washcoat porosity0.5
Pore diameter/nm0.56
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yu, S.; Zhang, J. Numerical Investigation on the Intraphase and Interphase Mass Transfer Limitations for NH3-SCR over Cu-ZSM-5. Processes 2021, 9, 1966. https://doi.org/10.3390/pr9111966

AMA Style

Yu S, Zhang J. Numerical Investigation on the Intraphase and Interphase Mass Transfer Limitations for NH3-SCR over Cu-ZSM-5. Processes. 2021; 9(11):1966. https://doi.org/10.3390/pr9111966

Chicago/Turabian Style

Yu, Shiyong, and Jichao Zhang. 2021. "Numerical Investigation on the Intraphase and Interphase Mass Transfer Limitations for NH3-SCR over Cu-ZSM-5" Processes 9, no. 11: 1966. https://doi.org/10.3390/pr9111966

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