Next Article in Journal
Eucalyptol: A Bio-Based Solvent for the Synthesis of O,S,N-Heterocycles. Application to Hiyama Coupling, Cyanation, and Multicomponent Reactions
Next Article in Special Issue
Magnesium Impregnated on NaX Zeolite Synthesized from Cogon Grass Silica for Fast Production of Fructose via Microwave-Assisted Catalytic Glucose Isomerization
Previous Article in Journal
Highly Efficient Heterogeneous Pd@POPs Catalyst for the N-Formylation of Amine and CO2
Previous Article in Special Issue
On the Support Effect and the Cr Promotion of Co Based Catalysts for the Acetic Acid Steam Reforming
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cu-IM-5 as the Catalyst for Selective Catalytic Reduction of NOx with NH3: Role of Cu Species and Reaction Mechanism

1
MOE Key Laboratory of Bioinorganic and Synthetic Chemistry, School of Chemistry, Sun Yat-sen University, Guangzhou 510275, China
2
State Key Laboratory of Catalytic Materials and Reaction Engineering, Research Institute of Petroleum Processing, SINOPEC, Beijing 100083, China
3
The ZeoMat Group, Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences, Qingdao 266101, China
*
Authors to whom correspondence should be addressed.
Catalysts 2021, 11(2), 221; https://doi.org/10.3390/catal11020221
Submission received: 20 January 2021 / Revised: 2 February 2021 / Accepted: 4 February 2021 / Published: 7 February 2021
(This article belongs to the Special Issue Catalysis with Ordered Porous Materials)

Abstract

:
The role of Cu species in Cu ion-exchanged IM-5 zeolite (Cu-IM-5) regarding the performance in selective catalytic reduction (SCR) of NOx with NH3 (NH3-SCR) and the reaction mechanism was studied. Based on H2 temperature-programmed reduction (H2-TPR) and electron paramagnetic resonance (EPR) results, Cu–O–Cu and isolated Cu species are suggested as main Cu species existing in Cu-IM-5 and are active for SCR reaction. Cu–O–Cu species show a good NH3-SCR activity at temperatures below 250 °C, whereas their NH3 oxidation activity at higher temperatures hinders the SCR performance. At low temperatures, NH4NO3 and NH4NO2 are key reaction intermediates. In situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) suggests a mixed Eley–Rideal (E–R) and Langmuir–Hinshelwood (L–H) mechanism over Cu-IM-5 at low temperatures.

Graphical Abstract

1. Introduction

Selective catalytic reduction (SCR) of nitrogen oxides (NOx) with NH3 is an effective way of eliminating environmentally harmful NOx in the exhaust gases from vehicles, ships, and electric plants. The biggest challenge of the research is to eliminate NOx from the oxygen-rich exhaust gas of diesel engines and to design practical diesel SCR catalysts [1]. Transition metal ion-exchanged zeolites (Cu, Fe, etc.) have been extensively studied and applied as commercial SCR catalysts in diesel-powered vehicles due to the excellent performance, i.e., high activity, high N2 selectivity, and hydrothermal stability [2].
Many zeolites with various framework types, such as MFI, BEA, CHA, AEI, ERI, KFI, AFX, DDR, RTH, SFW, LEV, LTA, RHO, and UFI, have been investigated as SCR catalysts [3,4,5,6,7,8]. Iwamoto et al. [9,10] first reported the catalytic activity of Cu-ZSM-5 for the decomposition of NO. Later, the catalyst became one of the most investigated zeolite materials for SCR. ZSM-5 (MFI topology) consists of an intersected 2D channel system, i.e., the straight channel along the b-axis and the sinusoidal channel along the a-axis. Both channels are delimited by 10-rings of TO4 (Si, Al) tetrahedral with pore openings 5.1 × 5.5 Å in size. Zeolite IM-5 (IMF topology) shows a similar 10-ring pore opening with ZSM-5, but a different 3D intersected channel system. IM-5 consists of two straight channels along the a-axis/c-axis, and a tortuous channel along the b-axis [11]. The zeolite is an excellent alternative to ZSM-5 in various reactions [12,13], especially for the NH3-SCR reaction. Vennestrøm et al. [14] found that IM-5 shows higher framework Al stability than ZSM-5 against hydrothermal treatments. However, the role of Cu species in SCR and the reaction mechanism were not fully understood yet.
In addition to the zeolite framework, Cu species are also studied intensively. As commonly accepted, isolated Cu2+ ions are the best catalytic active sites; CuAl2O4 species are inactive in the reaction; other species such as [Cu(OH)]+-Z and [Cu–O–Cu]2+ are still in debate [2,15,16,17,18]. Gao et al. [19] observed NH3 oxidation followed [Cu–O–Cu]2+ species formation under high levels of Cu ion-exchange degree. Recently, Liu et al. [20] demonstrated that highly dispersed CuOx (such as Cu–O–Cu) can also catalyze NOx reduction at a high temperature, i.e., 425 °C. [Cu(OH)]+-Z sites are more effective for NOx reduction at low temperatures but more selective for NH3 oxidation at high temperatures. Szanyi et al. [21,22] found that nitrosyl species bounded to copper are key intermediates during NH3-SCR. NO is firstly oxidized to NO2, then NO2 transforms to surface nitrates and nitrosyl species [23,24].
The purpose of this study is to illuminate Cu species in Cu-IM-5 and to illustrate the reaction mechanism toward NH3-SCR. In this study, Cu-IM-5 with different Cu contents were prepared through the ion-exchange method. The NH3-SCR performance was tested. Then, a combination of several characterization techniques is applied and correlated with the catalytic performance in NH3-SCR reaction, i.e., inductively coupled H2 temperature-programmed reduction (H2-TPR), NH3-temperature-programmed desorption (NH3-TPD), electron paramagnetic resonance (EPR), UV–vis-NIR, XPS, and in situ DRIFTS. In addition, the mechanism of SCR is investigated through in situ DRIFTS.

2. Results

2.1. Structure Characterization

A sample of zeolite IM-5 was chosen as the parent material, and a number of Cu-IM-5 catalysts with various Cu-loading amounts were prepared via ion-exchange in diluted Cu(NO3)2 solutions. Table 1 lists the elemental analysis of IM-5 and Cu-IM-5, and some other relevant characteristics. According to inductively coupled plasma atomic emission spectroscopy (ICP-AES) analysis, Si/Al ratios of IM-5 and ion-exchanged Cu-IM-5 samples remain practically unchanged (between 16.5 and 17.2). Cu contents of the ion-exchanged samples are 1.4 wt.%, 1.8 wt.%, and 2.3 wt.%, respectively, corresponding Cu exchange-rates are 0.56, 0.80, and 0.90. These samples are denoted according to their Cu contents, i.e., 1.4Cu-IM-5, 1.8Cu-IM-5, and 2.3Cu-IM-5, respectively.
Powder X-ray diffraction (PXRD) (Figure 1a) of IM-5 and Cu-IM-5 with different Cu contents show characteristic diffraction peaks of IMF topology, which are well fitted with the simulated pattern, suggesting the stability of the IM-5 framework under ion exchange conditions. Peaks that are characteristic of CuO (2θ = 35.6 and 38.8°) are absent in all Cu-IM-5 samples, implying that CuO crystals are not present even at higher ion-exchange rates. N2 adsorption-desorption isotherms of IM-5 and Cu-IM-5 samples (Figure 1b) are typical for the typical type I, which is characteristic of microporous materials. Calculated BET specific surface areas and pore volumes are listed in Table 1. IM-5 has a pore volume of 0.375 cm³/g with a BET surface area of 354.9 m2/g. After ion-exchanges with Cu, BET surface areas and pore volumes have decreased slightly with increasing Cu contents. As shown in Figure 1c,d, IM-5 consists of large particles, which are stacks of rod-like nanocrystals with lengths of ca. 300 nm. The morphology of 1.4Cu-IM-5, as an example for the ion-exchanged catalysts, is identical to the parent IM-5. The ion-exchange affects neither the crystal structure nor the morphology, as expected normally for zeolite materials.
As shown in Figure 2, the scanning transmission electron microscopy (STEM) image of 2.5Cu-IM-5 shows the typical morphology of IM-5, which is identical to that in Figure 1. The high dispersion of Cu species was further checked by EDS mapping.
29Si and 27Al MAS NMR spectra of IM-5 and Cu-IM-5 samples are illustrated in Figure 2. As shown in Figure 3a, 29Si resonances at −98, −101, −106, and −112 ppm are assigned to Q4(3), Q4(2), Q4(1), and Q4(0). Calculated Si/Al ratios by 29Si MAS NMR and ICP (Table 1) are in good agreement for IM-5 and Cu-IM-5 samples. In 27Al MAS NMR spectra (Figure 3b) of IM-5 and Cu-IM-5, only tetrahedral-coordinated Al are detected at δ = 58 ppm.

2.2. NH3-SCR Performance over Cu-IM-5

Figure 4a shows the NOx conversions during NH3-SCR with IM-5 and Cu-IM-5 samples. IM-5 show almost no activity for NOx reduction at low temperatures. For 1.4Cu-IM-5, NOx conversion shows a volcano-shaped line with increasing temperature. The maximum NOx conversion reaches 96.6% at 250 °C, then declines due to NH3 oxidation [14]. With increasing Cu contents, the NOx conversion increases slightly at low temperatures between 150 °C and 250 °C. However, in the high-temperature region (300–550 °C), NOx conversion declines with increasing Cu contents, which can be attributed to an increasing NH3 oxidation activity. It is worth noting that 1.4Cu-IM-5 exhibits higher NOx conversion than Cu-IM-5 from 200 °C to 550 °C. As shown in Figure 4b, high N2 selectivity was achieved overall Cu-IM-5 catalysts. N2 selectivity of 1.4Cu-IM-5 was 84.8% at 150 °C and then increases to 98% at 250 °C and starts to decline at 350 °C. The same trend of N2 selectivity against temperatures was found for 1.8Cu-IM-5 and 2.3Cu-IM-5 as well. Figure 4c shows NH3 oxidation on Cu-IM-5 from 250 °C to 550 °C. For NH3-SCR, NH3 oxidation is the primary side reaction, and NO is the main side product at high temperatures [25]. With increasing Cu contents, NH3 conversion rises, and NOx concentration increases. At high temperatures (>300 °C), there is insufficient NH3 to reduce NO, and NO is produced from NH3 oxidation. Hence, the apparent NOx conversion decreases at high temperatures. The slightly decreasing N2 selectivity with increasing Cu contents is also a consequence of the increasing NH3 oxidation.

2.3. Cu Species in Cu-IM-5

In order to evaluate the oxidation states of Cu species on the outer surfaces of Cu-IM-5, XPS analysis of Cu 2p core level was performed (Figure 5a). Peaks located at the binding energies of 933.8 ± 0.2 eV and 953.9 ± 0.2 eV are ascribed to Cu 2p3/2 and Cu 2p1/2, respectively. Shakeup peaks locate at 943 eV are essential characteristics of divalent Cu species [26]. Only one kind of peak at 933.8 eV assigned to Cu2+ species is observed, suggesting that all divalent copper species existing in Cu-IM-5 are similar. The intensity of this peak rises with increasing Cu contents.
EPR spectra reveal the number of isolated Cu ions (Figure 5b). Therefore, only isolated Cu2+ species can be detected, while other species such as Cu+ and CuOx clusters are inactive [27]. All samples show g = 2.05 and g// = 2.38, which are typical of isolated hydrated Cu2+ complexes in a distorted octahedral symmetry such as [Cu(H2O)6]2+ and [Cu(H2O)5(OH)]+ [28]. The intensity of g peak at high frequency decreases from 6.4 × 106 to 2.1 × 106 with increasing Cu contents. The decreasing intensity of g peak results from the loss of isolated Cu2+ ions, which aggregate and form Cu–O–Cu [28].
H2–TPR is to study the reducibility of Cu-IM-5 with various Cu contents. As shown in Figure 6, the curves are fitted to quantify different Cu species, and the results are shown in Table 2. Three kinds of peaks are clearly shown during 150–400 °C: the peaks centered at about 200 °C is attributed to the reduction of isolated [Cu(OH)]+ to Cu+; the peaks centered at about 240 °C is assigned to the reduction of divalent Cu in Cu–O–Cu type species to Cu+; the peak at 270 °C is assigned to the reduction of isolated Cu2+ species to Cu+; the peaks above 400 °C are assigned to the reduction of Cu+ to Cu0 [29,30]. For 1.4Cu-IM-5, the peak at 200 °C and 270 °C are observable, suggesting the presence of [Cu(OH)]+ and Cu2+ species. With Cu contents reach to 1.8 wt.%, a peak centered at 240 °C emerges, suggesting a formation of Cu–O–Cu species. The fraction of this peak increases when Cu contents further increase.
The acidity of Cu-IM-5 with different Cu contents is investigated using NH3-TPD (Figure 7). To quantify the amount of the NH3 desorbed from different acid sites, the desorption curves have been fitted, and the fraction of the peaks are shown in Table 3. Two peaks shown in the curve of IM-5—the low-temperature peak (peak α) centered at 200 °C is ascribed to desorption of NH3 from weak acid (WA) sites, and the high-temperature peak (peak γ) centered at 400 °C corresponds to the desorption of NH3 from strong acid (SA) sites. Cu-IM-5 samples exhibit a third peak (peak β) at 300 °C, which is attributed to the desorption of NH3 from Cu ions [31]. At the same time, the peak center of γ moves to low temperatures (374 °C), indicating that strong acid sites of the aluminosilicate framework become weaker due to sheltering effects of extra-framework Cu ions. The fraction of peak β increases with increasing Cu loadings.

2.4. Reaction Mechanisms

Taking 1.4Cu-IM-5 as an example, the catalyst is exposed to NO2, mixed with NO + O2 (Figure 8), and observed using DRIFTS. The bands in the range of 1650 cm−1 to 1500 cm−1 are attributed to nitrate or nitro species adsorbed on Cu sites [29,32]. Bands at 1624 cm−1, 1613 cm−1 (1597 cm−1), and 1573 cm−1 (1545 cm−1) are ascribed to adsorbed NO2 species, monodentate nitrates, bidentate nitrates, and nitrite adspecies [32,33], respectively. In NO2, the bands visibly increase with prolonging exposure time. Physically adsorbed NO2 is swept off through N2 for 30 min. Corresponding peaks decrease slightly. Exposed to mixed NO + O2, similar bands appear with lower intensity of IR bands. Wang et al. [32] reported that NO was first oxidized to NO2 on Cu-SAPO-34. It is widely recognized that NO oxidation is the key factor for the further formation of nitrate and nitrite species [34]. Compared to the IR bands in NO2 and NO + O2 flow, the relative intensity of the NO2 vibration band in the latter is stronger than that of the former, as a consequence that the existence of NO2 suppresses the oxidation of NO.
As shown in Figure 9, similar nitrate species are observed for Cu-IM-5 with different Cu contents after the exposure in NO + O2 flow. The intensity of the IR bands, especially the bands of NO2, increases with increasing Cu contents. It indicates that NO is easier oxidized on Cu-IM-5 with higher Cu content. Consequently, a higher amount of nitrates becomes observed by IR.
In situ DRIFTS using NH3 as a probe molecule is applied to test the acidity of Cu-IM-5. As illustrated in Figure 10, in situ DRIFTS are recorded over 1.4Cu-IM-5 that was exposed to 500 ppm NH3 flow and purged with N2 for 30 min. Bands at 3610, 3520, 3338, 3275, 3181, 1625, and 1480 cm−1 are observed, which are attributed to the adspecies listed in Table 3. In O–H stretch vibration region (3500~3800 cm−1), two negative bands located at 3610 cm−1 and 3520 cm−1 are assigned to Brønsted acid sites (Zeo–OH groups). They have been consumed by ammonia adsorption over Cu-IM-5 [35,36]. In N–H bending region (1350 cm−1~1700 cm−1), the band at 1484 cm−1 is attributed to asymmetric vibration of NH4+ adsorbed on Brønsted acid sites (δ(NH4+)), while the band at 1624 cm−1 is attributed to the N–H bonds of ammonia molecule coordinated with Cu sites [32,37]. In N–H stretching region, the bands at 3338 cm−1 and 3181 cm−1 can be assigned to ammonium ions, while the band at 3288 cm−1 can be assigned to coordinated ammonia molecule [31].
In situ DRIFTS study was carried out to follow the reaction between NO + O2 and pre-absorbed NH3 in Cu-IM-5, in order to investigate the variation of the adsorbed species on Cu-IM-5 during NH3-SCR reaction. As shown in Figure 11a, the intensity of the band at 1618 cm−1 declines at the beginning of the reaction. It almost disappears after 5 min, due to the consumption of NH3 species on Lewis acid (LA) sites. The band at 1480 cm−1 decreases slower and disappears until 15 min. The intensity of the band at 1618 cm−1 decreases faster than that of the band at 1480 cm−1. NH3 on LA reacts faster with NO + O2 than NH4+ on Brönsted acid (BA) sites. It is generally accepted that Cu ions (LA) are the active sites for SCR, while BA mainly existed as a reservoir of NH3 species [32,33,37]. NH4+ species on BA migrate to LA to participate NH3-SCR reactions [32].
New bands at 1625, 1613, 1595, 1575, and 1545 cm−1 appear after 5 min, while the peak at 1480 cm−1 can still be detected. The new bands correspond to formations of nitrates and nitrites.
Figure 11b shows the spectra that were taken on Cu-IM-5 with pre-adsorbed NOx during exposure to NH3. The intensity of bands at 1625, 1613, 1595, 1575, and 1545 cm−1 reveals the consumption of pre-adsorbed NOx species. The band at 1480 cm−1 that appears in 5 min suggests a formation of NH4+ on BA. The bands at 1613, 1595, 1575, and 1545 cm−1 still exist after 5 min. NH4NO3 may exist as an intermediate during SCR reaction [32]. The bands of NO2 and NO3 disappear completely after 30 min. Ultimately, bands at 1618 cm−1 and 1480 cm−1 appear.

3. Discussion

Cu contents play a significant role in the NH3-SCR performance of Cu-IM-5. NOx conversion increases slightly with increasing Cu contents at low temperatures (<200 °C) and decreases at higher temperatures (>200 °C). EPR results suggest the number of isolated Cu ions decreases with increasing Cu contents, while XPS results indicate an increased amount of Cu species on the outer surfaces of Cu-IM-5. A fraction of Cu ions agglomerated during calcination [4]. As indicated by H2–TPR, [Cu(OH)]+, Cu–O–Cu, and Cu2+ are the main species in Cu-IM-5. Cu–O–Cu species increase with rising Cu contents. Thus, the loss of NOx conversion at higher Cu contents is the consequence of higher amounts of Cu–O–Cu species. Similar to SSZ-13 [38], isolated Cu ions, i.e., [Cu(OH)]+ and Cu2+, are the active sites for NH3-SCR in Cu-IM-5. In situ DRIFTS results of NO + O2 adsorption demonstrates that more NO2 and NO3 species form on 2.3Cu-IM-5 (Figure 12). It is reasonable to conclude that NO is prone to oxidize on Cu–O–Cu species, which explains the high performance of 2.3Cu-IM-5 at low temperatures. Higher amounts of Cu–O–Cu species enhance the low-temperature activity of Cu-IM-5 (Table 4). On the contrary, at higher temperatures, NOx conversion decreases due to the higher NH3 oxidation.
NH4NO3 was observed as an important intermediate during the reactions both between adsorbed NH3 with NO + O2 and between adsorbed NOx with NH3. However, NH4NO2 was not observed, because it has a lower decomposition temperature of 80 °C [22]. NH4NO3 decomposes above 200 °C according to Reaction (1).
NH4NO3 → N2O + 2H2O
In situ DRIFTS was carried out at 150 °C, which is lower than the decomposition temperature. Moreover, N2 selectivity was very high at this temperature. NH4NO3 is consumed by reacting with NO according to reaction 2 [15].
NH4NO3 + NO → N2 + 2H2O + NO2
The freshly formed NO2 is to be adsorbed again and react further with NH3.

4. Materials and Methods

4.1. Synthesis of Cu-IM-5

As-synthesized IM-5 was provided by SINOPEC Research Institute of Petroleum Processing, Beijing, China. The material was calcined in the air following a temperature program with a plateau at 290 °C to 550 °C for 4 h. An ion-exchange in 1 L 0.25 M CH3COONH4 for 5 h at 80 °C was carried out. The obtained NH4–IM-5 was ion-exchanged to Cu-IM-5 in aqueous Cu(NO3)2 solution with different molarities (1 mM, 3 mM, and 30 mM), and calcined at 550 °C for 4 h. The obtained samples were denoted as 1.4Cu-IM-5, 1.8Cu-IM-5, and 2.3Cu-IM-5, corresponding to 1.4 wt.%, 1.8 wt.%, and 2.3 wt.% Cu contents, respectively.

4.2. Characterization of Cu-IM-5

Powder X-ray diffraction (PXRD) patterns were recorded on a Rigaku SmartLab diffractometer (Tokyo, Japan, Cu Kα radiation). A Hitachi SU8010 (Tokyo, Japan) scanning electron microscope (SEM) was used to observe crystal morphology and sizes. Elemental compositions were analyzed by ICP-AES on an Optima 8300 (PerkinElmer, Waltham, MA, USA). N2-adsorption isotherms were measured at 77 K with a Micromeritics ASAP 2020 Plus HD88 (Norcross, GA, USA) instrument. The t-plot method was used to calculate the total pore volumes (adsorption branch).
STEM image was carried out on JEM-ARM200P microscope (Tokyo, Japan) with 200 kV acceleration voltage. EDS analysis was performed using a Bruker XFlash 6T160 (Rheinstetten, Germany) apparatus.
29Si MAS NMR analyses were carried out on Bruker AVANCE III 600 spectrometer at a resonance frequency of 119.2 MHz. The spectra with high-power proton decoupling were recorded with a spinning rate of 10 kHz, a π/4 pulse length of 2.6 μs, and a recycle delay of 60 s. 27Al MAS NMR analyses were carried out at a resonance frequency of 156.4 MHz using a 4 mm HX double-resonance MAS probe at a sample spinning ratexl of 14 kHz. 27Al MAS NMR spectra were recorded by a small-flip angle technique with a pulse length of 0.68 μs (<π/12) and a 1 s recycle delay. The chemical shift of 27Al was referenced to 1 M aqueous Al(NO3)3.
Electron paramagnetic resonance (EPR) experiments were performed on A300-10-12 (Bruker) at the atmosphere.
H2 temperature-programmed reduction (TPR) was carried out on a homemade chemisorption analyzer equipped with a TCD. A total of 100 mg samples were put into a quartz tube and were pretreated at 350 °C in dry air for 1 h and cooled down to room temperature. H2–TPR was performed in 5% H2/N2 gas flow of 50 mL/min at a heating rate of 10 °C/min.
For NH3-temperature-programmed desorption (NH3-TPD), the catalyst was purged in N2 for 60 min at 350 °C and then exposed in an NH3/N2 flow for 30 min. Afterward, the catalyst was purged in He flow at 100 °C. NH3-TPD was measured in He of 50 mL/min from 100 to 800 °C with a ramp of 10 °C/min.
In situ DRIFTS experiments were carried out on a Nicolet IS20 spectrometer (Thermo Fisher, Waltham, MA, USA) with an MCT detector and a Harrick high-temperature reaction chamber with ZnSe windows. The sample was purged in 100 mL/min N2 flow at 500 °C for 30 min and then cooled down to the reaction temperature. The background spectra were recorded under this temperature. The spectra were measured in the range of 4000–650 cm−1 by accumulating 64 scans at a 4 cm−1 resolution. It is noted that in situ DRIFTS experiments were carried out without the existence of water.

4.3. NH3-SCR Test

SCR catalytic performance of Cu-IM-5 was tested on a fixed-bed quartz flow reactor (ID = 4 mm) at atmospheric pressure. 0.1 g catalysts (60–100 mesh) were used. The reaction feed was an N2-based gas mixture containing 500 ppm NO, 500 ppm NH3, and 5% O2. Gas hourly space velocity (GHSV) was fixed at 190,000 h−1. Reaction temperatures were from 150 °C to 550 °C. The concentrations of NOx were measured using IGS FT-IR (Thermo Fisher) equipped with a 2 m gas cell and an MCT detector with 4 cm−1 resolutions.
N 2   selectivity ( % )   =   ( NO   +   NH 3 ) inlet ( NO   +   NH 3 ) outlet   -   ( NO 2   +   2 N 2 O ) outlet ( NO + NH 3 ) inlet   -   ( NO + NH 3 ) outlet   ×   100 %
NO x   conversion ( % )   =   ( NO ) inlet - ( NO   +   NO 2   +   2 N 2 O ) outlet ( NO ) inlet   ×   100 %
The kinetic experiments were carried out with 50 mg catalyst (60–100 mesh) and with a volume hourly space velocity of 190,000 h−1.

5. Conclusions

The Cu-IM-5 catalysts exhibit high NH3-SCR performances, which show differences with increasing Cu contents. [Cu(OH)]+, Cu–O–Cu and isolated Cu2+ are the active sites for NH3-SCR. The amount of Cu–O–Cu species increases with increasing Cu contents. The low-temperature NH3-SCR activity increases with increasing Cu contents due to increasing Cu–O–Cu species. However, the Cu–O–Cu species are more active for NH3 oxidation at reaction temperatures above 350 °C. In situ DRIFTS suggests an L–H mechanism during NH3-SCR over Cu-IM-5 catalysts with NH4NOx (x = 2 or 3) as the intermediates.

Author Contributions

Conceptualization, G.F. and X.Y.; original draft preparation, G.F.; IM-5 preparation, J.C.; characterization, Y.L. and R.L.; writing—review and editing, X.Y. and J.J.; supervision, J.J.; funding acquisition, J.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China No. 21971259. Weifang Zhengxuan Rare Earth and Catalytic Materials Co., Ltd. is also thanked for financial supports.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chavannavar, P. Exhaust Aftertreatment System and Method. U.S. Patent No 9,132,386, 3 April 2014. [Google Scholar]
  2. Brandenberger, S.; Kröcher, O.; Tissler, A.; Althoff, R. The state of the art in selective catalytic reduction of NOx by ammonia using metal-exchanged zeolite catalysts. Catal. Rev. 2008, 50, 492–531. [Google Scholar] [CrossRef]
  3. Li, R.; Zhu, Y.; Zhang, Z.; Zhang, C.; Fu, G.; Yi, X.; Huang, Q.; Yang, F.; Liang, W.; Zheng, A.; et al. Remarkable performance of selective catalytic reduction of NOx by ammonia over copper-exchanged SSZ-52 catalysts. Appl. Catal. B Environ. 2021, 283, 9641. [Google Scholar] [CrossRef]
  4. Han, L.; Cai, S.; Gao, M.; Hasegawa, J.Y.; Wang, P.; Zhang, J.; Shi, L.; Zhang, D. Selective Catalytic Reduction of NOx with NH3 by Using Novel Catalysts: State of the Art and Future Prospects. Chem. Rev. 2019, 119, 10916–10976. [Google Scholar] [CrossRef]
  5. Zhang, R.; Liu, N.; Lei, Z.; Chen, B. Selective Transformation of Various Nitrogen-Containing Exhaust Gases toward N2 over Zeolite Catalysts. Chem. Rev. 2016, 116, 3658–3721. [Google Scholar] [CrossRef]
  6. Liu, B.; Chen, Z.; Huang, J.; Xia, Q.; Wu, Y.; Chen, H.; Fang, Y. Development of Iron Encapsulated Hollow Beta Zeolites for Ammonia Selective Catalytic Reduction. Ind. Eng. Chem. Res. 2019, 58, 2914–2923. [Google Scholar] [CrossRef]
  7. Liu, B.; Zheng, K.; Liao, Z.; Chen, P.; Chen, D.; Wu, Y.; Xia, Q.; Xi, H.; Dong, J. Fe-Encapsulated ZSM-5 Zeolite with Nanosheet-Assembled Structure for the Selective Catalytic Reduction of NOx with NH3. Ind. Eng. Chem. Res. 2020, 59, 8592–8600. [Google Scholar] [CrossRef]
  8. Liu, B.; Zhao, X.; Mao, W.; Chen, H.; Han, L.; Zhu, K.; Zhou, X. Pickering emulsion mediated crystallization of hierarchical zeolite SSZ-13 with enhanced NH3 selective catalytic reduction performance. Micropor. Mesopor. Mater. 2019, 285, 202–214. [Google Scholar] [CrossRef]
  9. Iwamoto, M.; Furukawa, H.; Mine, Y.; Uemura, F.; Mikuriya, S.I.; Kagawa, S.J. Copper(II) ion-exchanged ZSM-5 zeolites as highly active catalysts for direct and continuous decomposition of nitrogen monoxide. Chem. Soc. Chem. Commun. 1986, 1272–1273. [Google Scholar] [CrossRef]
  10. Li, J.; Chang, H.; Ma, L.; Hao, J.; Yang, R.T. Low-temperature selective catalytic reduction of NOx with NH3 over metal oxide and zeolite catalysts—A review. Catal. Today 2011, 175, 147–156. [Google Scholar] [CrossRef]
  11. Baerlocher, C.; Gramm, F.; Massüger, L.; McCusker, L.B.; He, Z.; Hovmöller, S.; Zou, X. Structure of the Polycrystalline Zeolite Catalyst IM-5 Solved by Enhanced Charge Flipping. Science 2007, 315, 1113–1116. [Google Scholar] [CrossRef] [PubMed]
  12. Tu, C.; Chen, J.; Li, W.; Wang, H.; Deng, K.; Vinokurov, V.A.; Huang, W. Hydrodeoxygenation of bio-derived anisole to cyclohexane over bi-functional IM-5 zeolite supported Ni catalysts. Sustain. Energy Fuels 2019, 3, 3462–3472. [Google Scholar] [CrossRef]
  13. Zhang, J.; Tian, F.; Chen, J.; Shi, Y.; Cao, H.; Ning, P.; Sun, S.; Xie, Y. Conversion of phenol to cyclohexane in the aqueous phase over Ni/zeolite bi-functional catalysts. Front. Chem. Sci. Eng. 2020. [Google Scholar] [CrossRef]
  14. Vennestrøm, P.N.R.; Janssens, T.V.W.; Kustov, A.; Grill, M.; Molina, P.A.; Lundegaard, L.F.; Tiruvalam, R.R.; Concepción, P.; Corma, A.J. Influence of lattice stability on hydrothermal deactivation of Cu-ZSM-5 and Cu-IM-5 zeolites for selective catalytic reduction of NOx by NH3. J. Catal. 2014, 309, 477–490. [Google Scholar] [CrossRef] [Green Version]
  15. Moretti, G.; Ferraris, G.; Fierro, G.; Jacono, M.L.; Morpurgo, S.; Faticanti, M.J. Dimeric Cu(I) species in Cu-ZSM-5 catalysts: The active sites for the NO decomposition. J. Catal. 2005, 232, 476–487. [Google Scholar] [CrossRef]
  16. Centi, G.; Perathoner, S. Nature of active species in copper-based catalysts and their chemistry of transformation of nitrogen oxides. Appl. Catal. Gen 1995, 132, 179–259. [Google Scholar] [CrossRef]
  17. Dzwigaj, S.; Janas, J.; Gurgul, J.; Socha, R.P.; Shishido, T.; Che, M. Do Cu(II) ions need Al atoms in their environment to make CuSiBEA active in the SCR of NO by ethanol or propane? A spectroscopy and catalysis study. Appl. Catal. B Environ. 2009, 85, 131–138. [Google Scholar] [CrossRef]
  18. Gao, F.; Peden, C.H.F. Recent Progress in Atomic-Level Understanding of Cu/SSZ-13 Selective Catalytic Reduction Catalysts. Catalysts 2018, 8, 140. [Google Scholar] [CrossRef] [Green Version]
  19. Gao, F.; Walter, E.D.; Washton, N.M.; Szanyi, J.; Peden, C.H.F. Synthesis and Evaluation of Cu-SAPO-34 Catalysts for Ammonia Selective Catalytic Reduction. 1. Aqueous Solution Ion Exchange. ACS Catal. 2013, 3, 2083–2093. [Google Scholar] [CrossRef]
  20. Liu, L.L.; Liu, L.; Cao, Y.; Wang, J.; Si, R.; Gao, F.; Dong, L. Controlling Dynamic Structural Transformation of Atomically Dispersed CuOx Species and Influence on Their Catalytic Performances. ACS Catal. 2019, 9, 9840–9851. [Google Scholar] [CrossRef]
  21. Kwak, J.H.; Lee, J.H.; Burton, S.D.; Lipton, A.S.; Peden, C.H.F.; Szanyi, J. A Common Intermediate for N2 Formation in Enzymes and Zeolites: Side-On Cu–Nitrosyl Complexes. Angew. Chem. Int. Ed. 2013, 52, 9985–9989. [Google Scholar] [CrossRef] [PubMed]
  22. Szanyi, J.; Kwak, J.H.; Zhu, H.; Peden, C.H.F. Characterization of Cu-SSZ-13 NH3 SCR catalysts: An in situ FTIR study. Phys. Chem. Chem. Phys. 2013, 15, 2368–2380. [Google Scholar] [CrossRef] [PubMed]
  23. Long, R.Q.; Yang, R.T. Reaction Mechanism of Selective Catalytic Reduction of NO with NH3 over Fe-ZSM-5 Catalyst. J. Catal. 2002, 207, 224–231. [Google Scholar] [CrossRef]
  24. Iwasaki, M.; Yamazaki, K.; Banno, K.; Shinjoh, H.J. Characterization of Fe/ZSM-5 DeNOx catalysts prepared by different methods: Relationships between active Fe sites and NH3-SCR performance. J. Catal. 2008, 260, 205–216. [Google Scholar] [CrossRef]
  25. Yu, T.; Hao, T.; Fan, D.; Wang, J.; Shen, M.; Li, W.J. Recent NH3-SCR Mechanism Research over Cu/SAPO-34 Catalyst. Phys. Chem. C 2014, 118, 6565–6575. [Google Scholar] [CrossRef]
  26. Chen, B.; Xu, R.; Zhang, R.; Liu, N. Economical way to synthesize SSZ-13 with abundant ion-exchanged Cu+ for an extraordinary performance in selective catalytic reduction (SCR) of NOx by ammonia. Environ. Sci. Technol. 2014, 48, 13909–13916. [Google Scholar] [CrossRef] [PubMed]
  27. Nanba, T.; Masukawa, S.; Ogata, A.; Uchisawa, J.; Obuchi, A. Active sites of Cu-ZSM-5 for the decomposition of acrylonitrile. Appl. Catal. B Environ. 2005, 61, 288–296. [Google Scholar] [CrossRef]
  28. Occhiuzzi, M.; Fierro, G.; Ferraris, G.; Moretti, G. Unusual Complete Reduction of Cu2+ Species in Cu-ZSM-5 Zeolites under Vacuum Treatment at High Temperature. Chem. Mater. 2012, 24, 2022–2031. [Google Scholar] [CrossRef]
  29. Peng, C.; Yan, R.; Peng, H.; Mi, Y.; Liang, J.; Liu, W.; Wang, X.; Song, G.; Wu, P.; Liu, F.J. One-pot synthesis of layered mesoporous ZSM-5 plus Cu ion-exchange: Enhanced NH3-SCR performance on Cu-ZSM-5 with hierarchical pore structures. Hazard. Mater. 2020, 385, 1593. [Google Scholar] [CrossRef]
  30. Shan, Y.; Du, J.; Yu, Y.; Shan, W.; Shi, X.; He, H. Precise control of post-treatment significantly increases hydrothermal stability of in-situ synthesized cu-zeolites for NH3-SCR reaction. Appl. Catal. B Environ. 2020, 266, 8655. [Google Scholar] [CrossRef]
  31. Gao, F.; Wang, Y.; Washton, N.M.; Kollár, M.; Szanyi, J.; Peden, C.H.F. Effects of Alkali and Alkaline Earth Cocations on the Activity and Hydrothermal Stability of Cu/SSZ-13 NH3-SCR Catalysts. ACS Catal. 2015, 5, 6780–6791. [Google Scholar] [CrossRef]
  32. Wang, D.; Zhang, L.; Kamasamudram, K.; Epling, W.S. In Situ-DRIFTS Study of Selective Catalytic Reduction of NOx by NH3 over Cu-Exchanged SAPO-34. ACS Catal. 2013, 3, 871–881. [Google Scholar] [CrossRef]
  33. Ma, L.; Cheng, Y.; Cavataio, G.; McCabe, R.W.; Fu, L.; Li, J. In situ DRIFTS and temperature-programmed technology study on NH3-SCR of NOx over Cu-SSZ-13 and Cu-SAPO-34 catalysts. Appl. Catal. B Environ. 2014, 156–157, 428–437. [Google Scholar] [CrossRef]
  34. Olsson, L.; Sjövall, H.; Blint, R.J. Detailed kinetic modeling of NOx adsorption and NO oxidation over Cu-ZSM-5. Appl. Catal. B Environ. 2009, 87, 200–210. [Google Scholar] [CrossRef]
  35. Zhu, H.; Kwak, J.H.; Peden, C.; Szanyi, J. In situ DRIFTS-MS studies on the oxidation of adsorbed NH3 by NOx over a Cu-SSZ-13 zeolite. Catal. Today 2013, 205, 16–23. [Google Scholar] [CrossRef]
  36. Zhang, T.; Shi, J.; Liu, J.; Wang, D.; Zhao, Z.; Cheng, K.; Li, J. Enhanced hydrothermal stability of Cu-ZSM-5 catalyst via surface modification in the selective catalytic reduction of NO with NH3. Appl. Surf. Sci. 2016, 375, 186–195. [Google Scholar] [CrossRef]
  37. Sjövall, H.; Fridell, E.; Blint, R.J.; Olsson, L. Identification of adsorbed species on Cu-ZSM-5 under NH3 SCR conditions. Top. Catal. 2007, 42, 113–117. [Google Scholar] [CrossRef]
  38. Song, J.; Wang, Y.; Walter, E.D.; Washton, N.M.; Mei, D.; Kovarik, L.; Engelhard, M.H.; Prodinger, S.; Wang, Y.; Peden, C.H.F.; et al. Toward Rational Design of Cu/SSZ-13 Selective Catalytic Reduction Catalysts: Implications from Atomic-Level Understanding of Hydrothermal Stability. ACS Catal. 2017, 7, 8214–8227. [Google Scholar] [CrossRef]
Figure 1. (a) Powder X-ray diffraction (PXRD) patterns, (b) N2 adsorption and desorption curves, (c) SEM images of IM-5, and (d) 1.4Cu-IM-5.
Figure 1. (a) Powder X-ray diffraction (PXRD) patterns, (b) N2 adsorption and desorption curves, (c) SEM images of IM-5, and (d) 1.4Cu-IM-5.
Catalysts 11 00221 g001
Figure 2. STEM images (a,b; scale bar corresponds 50 nm); and EDS mapping (bf; scale bar corresponds 100 nm) of 2.3Cu-IM-5.
Figure 2. STEM images (a,b; scale bar corresponds 50 nm); and EDS mapping (bf; scale bar corresponds 100 nm) of 2.3Cu-IM-5.
Catalysts 11 00221 g002
Figure 3. (a) Si and (b) 27Al MAS NMR spectra of Cu-IM-5 with different Cu loadings.
Figure 3. (a) Si and (b) 27Al MAS NMR spectra of Cu-IM-5 with different Cu loadings.
Catalysts 11 00221 g003
Figure 4. (a) NOx conversion, (b) N2 selectivity, and (c) NH3 oxidation over IM-5 and Cu-IM-5 with different Cu contents.
Figure 4. (a) NOx conversion, (b) N2 selectivity, and (c) NH3 oxidation over IM-5 and Cu-IM-5 with different Cu contents.
Catalysts 11 00221 g004
Figure 5. (a) XPS (b) and electron paramagnetic resonance (EPR) of Cu-IM-5 with different Cu loadings.
Figure 5. (a) XPS (b) and electron paramagnetic resonance (EPR) of Cu-IM-5 with different Cu loadings.
Catalysts 11 00221 g005
Figure 6. Temperature-programmed reduction (TPR) curve of Cu-IM-5 with different Cu loadings: (a) 1.4Cu-IM-5, (b) 1.8Cu-IM-5, and (c) 2.3Cu-IM-5.
Figure 6. Temperature-programmed reduction (TPR) curve of Cu-IM-5 with different Cu loadings: (a) 1.4Cu-IM-5, (b) 1.8Cu-IM-5, and (c) 2.3Cu-IM-5.
Catalysts 11 00221 g006
Figure 7. NH3-temperature programmed desorption (NH3-TPD) of Cu-IM-5 with different Cu loadings: (A) IM-5, (B) 1.4Cu-IM-5, (C) 1.8Cu-IM-5, and (D) 2.3Cu-IM-5.
Figure 7. NH3-temperature programmed desorption (NH3-TPD) of Cu-IM-5 with different Cu loadings: (A) IM-5, (B) 1.4Cu-IM-5, (C) 1.8Cu-IM-5, and (D) 2.3Cu-IM-5.
Catalysts 11 00221 g007
Figure 8. In situ DRIFTS of 1.4Cu-IM-5 exposed in (a) 500 ppm NO2 and (b) 500 ppm NO + 5% O2 followed by purging in N2 for 30 min at 100 °C.
Figure 8. In situ DRIFTS of 1.4Cu-IM-5 exposed in (a) 500 ppm NO2 and (b) 500 ppm NO + 5% O2 followed by purging in N2 for 30 min at 100 °C.
Catalysts 11 00221 g008
Figure 9. In situ DRIFTS of Cu-IM-5 with different Cu loadings exposed in 500 ppm NO + 5% O2.
Figure 9. In situ DRIFTS of Cu-IM-5 with different Cu loadings exposed in 500 ppm NO + 5% O2.
Catalysts 11 00221 g009
Figure 10. In situ DRIFTS spectra of Cu-IM-5 exposed in (a) 500 ppm NH3 and (b) purged with N2.
Figure 10. In situ DRIFTS spectra of Cu-IM-5 exposed in (a) 500 ppm NH3 and (b) purged with N2.
Catalysts 11 00221 g010
Figure 11. In situ DRIFTS of (a) reaction between 500 ppm NO + 5% O2 and pre-absorbed NH3 on 1.4Cu-IM-5 catalysts and (b) reaction between 500 ppm NH3 and pre-absorbed NOx on 1.4Cu-IM-5.
Figure 11. In situ DRIFTS of (a) reaction between 500 ppm NO + 5% O2 and pre-absorbed NH3 on 1.4Cu-IM-5 catalysts and (b) reaction between 500 ppm NH3 and pre-absorbed NOx on 1.4Cu-IM-5.
Catalysts 11 00221 g011
Figure 12. The framework of Cu-IM-5 and adsorbed NOx on different Cu species.
Figure 12. The framework of Cu-IM-5 and adsorbed NOx on different Cu species.
Catalysts 11 00221 g012
Table 1. Element analysis and physical properties of the relevant samples.
Table 1. Element analysis and physical properties of the relevant samples.
SampleSi/Al
(ICP)
Framework
Si/Al 1
Cu Contents
wt.%
Cu/2AlEPR Signal Intensity
(g)
SBET
m2/g
Pore Volume
cm³/g
IM-516.513.9000354.90.375
1.4Cu-IM-516.812.81.40.566.4 × 106347.70.370
1.8Cu-IM-517.212.51.80.803.2 × 106336.30.358
2.3Cu-IM-516.712.32.30.902.1 × 106328.90.345
1 Calculated by 29Si MAS NMR.
Table 2. H2–TPR peak fractions of different Cu species.
Table 2. H2–TPR peak fractions of different Cu species.
Samples[Cu(OH)]+Cu–O–Cu Cu2+
1.4Cu-IM-515.223.761.1
1.8Cu-IM-513.049.737.3
2.3Cu-IM-512.156.131.8
Table 3. Fractions of different desorption peaks in NH3-TPD curve.
Table 3. Fractions of different desorption peaks in NH3-TPD curve.
SampleFraction of Different Peaks
Weak AcidCu SpeciesStrong Acid
IM-50.407-0.593
1.4Cu-IM-50.4460.1720.381
1.8Cu-IM-50.4830.1920.324
Table 4. Adsorbed species over 1.4Cu-IM-5 exposed in NH3 (Figure 10) or NO+O2 (Figure 8), and appear (a) and disappear (d) time during reaction (Figure 11).
Table 4. Adsorbed species over 1.4Cu-IM-5 exposed in NH3 (Figure 10) or NO+O2 (Figure 8), and appear (a) and disappear (d) time during reaction (Figure 11).
Wave Number (cm−1)GroupAppear (a) and Disappear (d) Time (min)
Figure 11aFigure 11b
3610Zeo–OH group--
3275N–H stretching of NH4+ on BA--
3181 and 3338N–H stretching of NH3 on LA--
1618N–H bending of NH3 on LA5 (d)8 (a)
1480N–H bending of NH4+ on BA15 (d)5 (a)
1625NO25 (a)10 (d)
1615 and 1597monodentate nitrates5 (a)10 (d)
1575bidentate nitrates5 (a)15 (d)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/catal11020221

AMA Style

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(2):221. https://doi.org/10.3390/catal11020221

Chicago/Turabian Style

Fu, Guangying, Junwen Chen, Yuqian Liang, Rui Li, Xiaobo Yang, and Jiuxing Jiang. 2021. "Cu-IM-5 as the Catalyst for Selective Catalytic Reduction of NOx with NH3: Role of Cu Species and Reaction Mechanism" Catalysts 11, no. 2: 221. https://doi.org/10.3390/catal11020221

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