Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2024 May 31;4(4):405–421. doi: 10.1021/acsengineeringau.4c00004

Effect of SO2 and SO3 Exposure to Cu-CHA on Surface Nitrate and N2O Formation for NH3–SCR

Joonsoo Han †,*, Joachim D Bjerregaard , Henrik Grönbeck , Derek Creaser , Louise Olsson †,*
PMCID: PMC11342297  PMID: 39185390

Abstract

graphic file with name eg4c00004_0008.jpg

We report effects of SO2 and SO3 exposure on ammonium nitrate (AN) and N2O formation in Cu-CHA used for NH3–SCR. First-principles calculations and several characterizations (ICP, BET, XRD, UV–vis–DRS) were applied to characterize the Cu-CHA material and speciation of sulfur species. The first-principles calculations demonstrate that the SO2 exposure results in both (bi)sulfite and (bi)sulfate whereas the SO3 exposure yields only (bi)sulfate. Furthermore, SOx adsorption on framework-bound dicopper species is shown to be favored with respect to adsorption onto framework-bound monocopper species. Temperature-programmed reduction with H2 shows two clear reduction states and larger sulfur uptake for the SO3-exposed Cu-CHA compared to the SO2-exposed counterpart. Temperature-programmed desorption of formed ammonium nitrate (AN) highlights a significant decrease in nitrate storage due to sulfur species interacting with copper sites in the form of ammonium/copper (bi)bisulfite/sulfate. Especially, highly stable sulfur species from SO3 exposure influence the NO2–SCR chemistry by decreasing the N2O selectivity during NH3–SCR whereas an increased N2O selectivity was observed for the SO2-exposed Cu-CHA sample. This study provides fundamental insights into how SO2 and SO3 affect the N2O formation during ammonium nitrate decomposition in NH3–SCR applications, which is a very important topic for practical applications.

Keywords: SO2/SO3-exposure, Cu-CHA, TPD (temperature-programmed desorption), AN (ammonium nitrate), N2O

1. Introduction

NH3-assisted selective catalytic reduction (SCR) has been extensively applied for NOx emission control in stationary and mobile applications (e.g., power plants, automobiles, and ships). Copper-exchanged chabazite (Cu-CHA), composed of a small pore entrance (eight-membered rings, 3.8 × 3.8 Å) and double six-membered rings, has been successfully commercialized in exhaust after treatment systems (EATS) for diesel combustion thanks to its high activity and hydrothermal stability at low and high temperatures, respectively.13 Nitrous oxide (N2O) emissions, which may be formed via side reactions during NOx reduction, must be regulated because of their ∼300 times stronger greenhouse gas potential than carbon dioxide (CO2) and depletion of the stratospheric ozone.4,5 As a result, there is a great deal of interest in understanding the fundamental mechanisms of N2O formation during NH3–SCR of NOx to improve N2 selectivity and minimize N2O emissions.

Utilization of biomass in fuels is indispensable to achieve sustainability goals and reduce fossil fuel usage, resulting in lower fossil carbon dioxide (CO2) emissions.6 However, combustion of petrol, diesel, and biofuels and usage of engine lubricants result in exhaust containing common poisons such as alkali metals, sulfur, phosphorus, and ash, causing contamination of the catalyst and lowering overall deNOx performance in EATS.7,8 Among the contamination sources, sulfur has been measured to have severe effects on the activity of the SCR catalysts.916 Sulfur is present in the form of sulfur oxides (SOx) after combustion of the fuels and originates mainly from the fuel. SO2-exposed Cu-CHA shows reduced deNOx performance compared to degreened Cu-CHA.11,12 SO3-exposed Cu-CHA, however, shows critically reduced deNOx performance and involves larger sulfur storage compared to SO2-exposed Cu-CHA.9,17 Thus, there is a strong demand to find strategies to reduce the effects of sulfur to maintain the catalytic activity during the lifetime of the catalyst.7 In modern diesel EATS, a diesel oxidation catalyst (DOC) plays an important role to increase the NO2 content, which promotes NOX reduction and passive regeneration of the SCR and diesel particulate filter (DPF), respectively.18 Unfortunately, SO2 is mainly oxidized via the DOC and coated DPF, resulting in SO3 and H2SO4 (formed with water via SO3+H2O → H2SO4) in the exhaust gas mixture.18,19

Although Europe follows the fuel standard (EN590) limiting the sulfur concentration to a maximum of 10 ppm(w), a substantial amount of sulfur (around 2.5 kg of S) could reach the catalysts in the EATS for heavy-duty applications during its lifespan.8,20 Periodic regeneration of the DPF (also referred to as active regeneration) can cause temperatures reaching over 650 °C in the filter, which means that also the downstream SCR catalyst is exposed to these high temperatures.21 During the high temperature, the accumulated surface sulfur species can leave the surface of the oxidation catalysts (DOC/DPF) whereupon the resulting sulfur oxides (SO2, SO3, and H2SO4) reach the SCR catalyst. According to the SOx-TPD with an EATS configuration (SCR+DOC+DPF+SCR+ASC), the DOC is the main contributor of the SO3/H2SO4 generation.22 With a sulfated SCR catalyst, the adsorbed sulfur species cause CuO formation during the desulfation event. This desulfation event can cause critical loss of the SCR active copper ions to inactive CuxOy species without severe zeolite structure deterioration.23

NH3–SCR reactions undergo different reaction schemes depending on the NO and NO2 ratio in the exhaust gas mixture.24,25 With only NO present, the NH3–SCR reaction includes O2 in the reaction stoichiometry and the reaction follows the so-called standard SCR reaction.24 The SCR catalyst can be exposed to NO2, which can be formed via an oxidation catalyst located upstream of the SCR catalyst. When using equimolar NO/NO2 amounts, the NOx reduction rate is enhanced compared to standard SCR, in the so-called fast SCR reaction.25 However, when the NO2 content is high, the slow NO2 SCR reaction occurs, which reduces the reaction rate even when compared to the standard SCR reaction.25

Standard SCR: 4NO + 4NH3 + O2 → 4N2 + 6H2O

Fast SCR: NO + NO2 + 2NH3 → 2N2 + 3H2O

NO2 SCR (slow SCR): 3NO2 + 4NH3 → 3.5N2 + 6H2O

A variety of Cu+/Cu2+ active sites have been identified depending on the temperature and gas composition in Cu-CHA. At a low temperature, H2O or NH3 solvation of the isolated copper sites causes framework-detached mobile Cu species such as Cu2+(H2O)x, Cu2+(OH)(H2O)x, Cu2+2O2(NH3)4, and Cu+(NH3)2.26 At a high temperature, H2O or NH3 desorption from these mobile Cu-monomer or Cu-dimer species produces framework-bound Cu-monomer (ZCu, ZCu(OH), Z2Cu)2629 and oxygen- or hydrogen-containing dimer (Z2Cu2O, Z2Cu2O2, Z2Cu2OxHy) species.27,3032 The Z denotes a negatively charged framework, and its subscription represents the number of the site. However, under standard SCR conditions at low temperatures (below 250 °C) in the presence of ammonia, the copper species are mobile and the NH3-solvated copper dimers ([Cu2(NH3)4O2]2+) and monomers ([Cu(NH3)2]+) are considered to be key intermediates in the SCR cycle.2,24,26,33

Bimodal N2O release is measured with a peak at low (below 300 °C) and high (above 300 °C) temperatures during exposure of Cu-CHA in standard SCR conditions.3436 According to the standard SCR mechanism at low temperatures, recent experimental studies have shown that N2O is only formed in the reduction half cycle (RHC) during the catalytic turnover of the standard SCR cycle.37 Ab initio calculations describing H2NNO decomposition over [Cu2(NH3)4OOH]2+ complexes explain not only N2O formation during the RHC but also the trend of increasing N2O formation as a function of Cu content.36 Low NO oxidation rates compared to NH3–SCR rates on Cu-CHA have been reported by Paolucci et al.,28 suggesting that NO oxidation is irrelevant to the standard SCR at low temperatures. In addition, the N2O formation rate is an order of magnitude higher than the NO2 formation rate.37 A pathway where N2O is formed via ammonium nitrate (AN) is, thus, not likely to contribute to the N2O formation during low-temperature standard SCR. In contrast to the standard SCR, the extent of N2O formation is significantly increased when the NO2 fraction is increased in the gas mixture5 and the 2+ oxidation state of Cu ions is predominantly observed from in situ XAS over Cu-CHA under fast and NO2 SCR conditions,38,39 suggesting that different reaction schemes are involved with respect to the fast and NO2 SCR conditions. Kubota et al.40 and Liu et al.41 demonstrated that AN is an intermediate species for N2O under fast/NO2 SCR conditions. Liu et al.41 proposed that the fast SCR follows a dual-site mechanism where derived intermediate species (HONO and H2NNO) are decomposed at Brønsted sites in a similar manner as the standard SCR scheme suggested by Chen et al.42 Han et al.5 proposed that the CHA structure is favorable to form surface nitrates due to promotion of the NO2 disproportionation within the CHA cage compared to the MFI and BEA structures, and thereby more AN is observed in the CHA structure, followed by MFI and BEA.

After SO2 exposure, only small or negligible amounts of stored sulfur has been observed in H forms of zeolites upon sulfur exposure,4345 indicating that SO2 storage mainly requires the presence of Cu ions. Physical and chemical poisoning is reported as the sulfur deactivation mechanism in Cu zeolites. SO2 and NH3 exposure to Cu-CHA shows decreasing micropore volume with increasing S content, suggesting that pore blocking occurs by ammonia interacting with derived sulfur species.46 In comparison, only SO2 exposure shows no considerable micropore volume decrease with S content.46 Temperature-programmed desorption (TPD) tests with an SO2+SCR gas composition gives bimodal SO2 peaks at low (∼400 °C) and high (↑600 °C) temperatures where ammonium sulfate and Cu sulfate are associated, respectively.10,45,4749 After regeneration, ammonium sulfate is decomposed and thus it causes a reversible deactivation.9,47 However, chemically interacting sulfur species at Cu sites (i.e., ZCuHSO4) or Al sites (i.e., Al2(SO4)3) are not fully regenerated and these derived sulfur species are highly stable.9,50 Furthermore, a fairly high temperature (over 700 °C) is necessary to fully desulfate these highly stable sulfur species over Cu-CHA.9,50 A highly thermal regeneration event can result in a hydrothermal aging (HTA) effect that converts the Cu ions to inactive CuXOy and CuAlOx species as well as CHA structure degradation.51 Therefore, high-temperature regeneration can result in irreversible thermal deactivation.

The current mechanistic understanding of the deactivation of SCR catalysts resulting from sulfur is mainly focused on SO2 exposure of Cu-CHA. There is a strong consensus that SO2 readily interacts with ZCuOH and that the interaction with Z2Cu is weaker.45,48,5254 In situ XAS shows that SO2 interaction with Cu ions, especially NH3-solvated mobile Cu dimers ([Cu2(NH3)4O2]2+), is particularly sensitive to SO2 as compared to other-state Cu ions such as framework-interacting Cu+/Cu2+ (ZCu/Z2Cu) or NH3-solvated Cu monomers ([Cu(NH3)2]+).55 Another XAS study shows that SO3 exposure does not change the local structure of Cu sites within the zeolite framework (i.e., Cu–O coordination number and the Cu–Cu distance, etc.), indicating that Cu sites remain highly dispersed after sulfation as well as following desulfation.56 First-principles calculations performed by Bjerregaard et al.57 agree with the experimental observations and suggest that the standard SCR deactivation at low temperatures is due to hindered intercage diffusion by ammonium (bi)sulfate species. Thereby, the pairing of Cu-monomer complexes ([Cu(NH3)2]+) via O2 activation is hindered, which is a key intermediate step of the oxidation half cycle (OHC) during standard SCR at low temperatures.34 The mechanism proposed by Bjerregaard et al.57 rationalizes the experimental findings, where the NO conversion and the N2O formation are reduced upon SO2 exposure of Cu-CHA during the standard SCR at low temperature.12,58,59 Mesilov et al.12 reported that NOx conversion and N2O selectivity are decreased for the SO2-exposed Cu-CHA (Si/Al ≈ 7, 3 wt % Cu) during the standard, fast, and NO2 SCR reactions. After regeneration, the NOx conversion and N2O selectivity were not fully regenerated for the three overall SCR reactions but the N2O selectivity was noticeably increased for the NO2 SCR compared to the degreened state. A comparative study between the Cu-CHA aged in EATS (270,000–710,000 miles) and HTA+SOx-treated Cu-CHA suggests that sulfur species assists CuO formation without a clear structural damage during desulfation.23 In contrast to the SO2 exposure, SO3 exposure to Cu-CHA results in larger sulfur uptake and more severe irreversible deactivation.9

For industrial applications, understanding the fundamentals of SOX chemistry in Cu-CHA as part of an NH3 SCR system is critically important to establish an effective desulfation strategy that maintains catalytic activity throughout the usage lifetime of the SCR catalyst. Limited studies are reported with respect to the SO3 exposure compared to the SO2 exposure effects on Cu-CHA in terms of its deNOx performance and N2O release. There are some studies that have reported the SO2 exposure effect on the N2O formation.12,37,58,60 However, to the best of our knowledge, there is no comparative study of the evaluation of the AN and N2O formation for the SO2- and SO3-exposed Cu-CHA, which is the objective of the current study.

2. Experimental and Computational Methods

2.1. Catalyst Synthesis and Monolith Sample Preparation

Copper-exchanged SSZ-13 (CHA) was prepared by following the hydrothermal synthesis method to acquire a Si to Al molar ratio of ≈15 and ca. 1 wt % Cu. Specifically, 1 wt % Cu was set to favor the exclusive presence of ion-exchanged Cu ions inside CHA cage and avoid possible formation of CuO particles. The acquired Cu/SSZ-13 powders were washcoated on honeycomb monolith substrates (cordierite, 400 cpsi, 15 mm × 20 mm). The specific synthesis and washcoating procedure are described in the Supporting Information (S1. Catalyst synthesis and monolith preparation). The resulting fresh Cu/SSZ-13 powder and monoliths were degreened under standard SCR conditions (400 ppm of NH3/NO + 10% O2 + 5% H2O + Ar as a balance) at 750 °C for 5 h prior to powder sample characterizations and experiments in a synthetic gas bench reactor.

2.2. Synthetic Gas Bench (SGB) Reactor Setup

An SGB reactor was used to perform degreening of fresh catalyst powders, SOx treatment, and AN temperature-programmed desorption (AN-TPD). Mass flow controllers (MFCs, Bronkhorst) and a controlled evaporator and mixing (CEM, Bronkhorst) system were equipped to provide the required gases (Ar, O2, NH3, NO, NO2, N2O, N2, H2O, and SO2). The gas lines were carefully insulated and heated to 191 °C to avoid water condensation and deposition of solids such as AN. Ar was used as an inert balance for all the experiments, and the total flow was 1.2 NL·min–1. The required gas mixture was fed to a horizontal quartz tube (inner diameter: 16 mm). The horizontal quartz tube was wrapped with a heating coil and insulated to heat up and control the reactor temperature. The prepared sample was located inside the quartz tube during the pretreatment and experiments.

2.3. SO2 and SO3 Exposure to Degreened Cu-CHA

The prepared monoliths were pretreated in the SGB reactor with SO2 and a SO2+SO3 mixture, which will be referred to as SO2 and SO3 exposure treatments. First, the monolith was degreened under standard SCR conditions (400 ppm of NH3, 400 ppm of NO, 10% O2, and 5% H2O) at 750 °C for 5 h, following the SO2 or SO3 treatment sequence, consisting of the following steps: first step: 30 ppm of SO2 + base feed (10% O2 + 5% H2O + Ar) for 1 h; second step: 400 ppm of NH3 + base feed for 2 h; third step: 30 ppm of SO2 + base feed for 1 h. The reason for applying the ammonia in a separate step between the sulfur steps and not simultaneously with SOx is that feeding SO3 and NH3 simultaneously causes large ammonium sulfate deposits in the reactor, which could block the lines, even when heating the lines thoroughly. For the SO2+SO3 treatment, an external device connected upstream from the SGB reactor was used to generate SO3. Pt/Al2O3 (7.5 wt % Pt) was used as an oxidation catalyst in the SO3 generator. The specific procedures are described in the Supporting Information (S2. SO3 calibration and SOx-exposure of degreened monoliths). This provided the capability to introduce 24 ppm of SO3 + 6 ppm of SO2 with base feed to the monolith instead of 30 ppm of SO2, as illustrated in Figure S2. Meanwhile, the temperature of the monolith was maintained at 400 °C during the SOx treatment to mimic the SOx exposure environment to the SCR catalyst during periodic regeneration of the DPF. Note that for simplification from here on, the SO2 and SO2+SO3 treatments shall be referred to as SO2 and SO3 exposure, respectively.

2.4. Catalyst Characterization

Elemental analysis was performed with powder forms of the degreened and SO2- and SO3-exposed samples. The SOx-exposed samples’ washcoats were carefully scraped from the SO2- and SO3-exposed monoliths. Afterward, the SOx-exposed washcoat samples were used for the characterizations. However, for diffuse reflectance infrared Fourier transform spectroscopy (ICP) and diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) results of the sulfated samples, crushed monoliths were used.

2.4.1. Catalyst Powder Sample Characterization

Elemental compositions of the fresh catalyst powders were analyzed using inductively coupled plasma sector field mass spectrometry (ICP-SFMS) at ALS Scandinavia AB. For sulfated samples, washcoated sulfated monoliths were crushed well into a fine powder form to acquire the S/Cu molar ratio; afterward, the sulfur content of the sulfated samples was determined.

Powder X-ray diffraction (XRD) was performed with degreened and SO2- and SO3-exposed samples to confirm the crystalline structure of CHA. A Siemens D5000 diffractometer operating at 40 kV, using the Kα1 radiation of a Cu anode as X-ray source (λ = 1.54060 Å), was used for measuring the X-ray diffractogram from the powder samples.

N2 physisorption was performed to measure specific surface area using the Brunauer–Emmett–Teller (BET) method of the degreed catalyst powder. A Micromeritics TriStar 3000 instrument was used to measure the N2 adsorption/desorption isotherms at 77 K. Prior to the measurements, the H and Cu forms of the degreened catalyst powders were degassed at 225 °C for 8 h.

UV–visible diffuse reflectance spectroscopy (UV–vis-DRS) was performed to investigate the formation of copper and ammonium sulfite/sulfate species. A Lambda 365 (PerkinElmer) double-beam UV visible spectrophotometer was used to acquire the UV spectrum. The system is composed of two different lamps (D2 and tungsten lamp) for measurement within the UV and visible-light regions. The lamp change occurred at 380 nm during the measurement. The light is split into two beams before it interacts with the sample. One of the beams is used as a reference, and the other beam reaches the sample. Two detectors (photodiodes) measured the light intensity, which came from the reference and sample beams simultaneously. Detector and reflectance sphere (sphere diameter: 60 mm, sample port aperture: 12.5 mm, sphere coating: Barium sulfate) modules were installed. Thereafter, a light trapper and a white standard were used for the calibration of UV intensity. Finally, the prepared catalyst powder was transferred to a powder cell for measuring the powder sample. The sample was scanned from 1100 to 210 at 5 nm slit width and 480 nm·min–1 scanning speed.

2.4.2. H2 Temperature-Programmed Reduction (H2-TPR)

H2-TPR was carried out with a differential scanning calorimeter (Sensys DSC calorimeter from Setaram) to investigate the reduction properties of the degreened and SO2- and SO3-exposed catalyst powders. The powders were carefully sieved to 180–250 μm fraction. Degreened (ca. 50 mg) or SOx-exposed (ca. 26 mg) samples were loaded into a quartz tube located inside the DSC, and 0.2% H2/Ar was continuously fed at 20 N mL·min–1 during the experiment. Initially, the temperature was set at 25 °C for 20 min and then was ramped up to 800 °C at 10 °C·min–1; thereafter, 800 °C was maintained for 20 min. Note that there was no pretreatment done prior to H2-TPR. In the meantime, the outlet gases from the DSC were monitored with mass spectrometry (Hidden HPR-20 QUI MS) for H2 (m/e = 2), NH3 (15), H2O (18), Ar (20), O2 (32), H2S (34), SO2 (64), SO3 (80), and H2SO4 (98).

2.4.3. Ammonium Nitrate Temperature-Programmed Desorption (AN-TPD)

AN-TPD was performed to investigate the effect of SO2 and SO3 on AN and N2O formation in the SGB reactor. 200 ppm of NH3/NO2 + 5% H2O + Ar was fed to the monolith for 90 min at 150 °C, followed by purging with 5% H2O+Ar for 90 min. Thereafter, the temperature was increased to 550 °C at 10 °C·min–1 and then maintained at 550 °C for 20 min while the monolith was exposed to 5% H2O+Ar. Note that no pretreatment was done prior to AN-TPD for the SO2- and SO3-exposed samples, to avoid the removal of sulfur species. However, the degreed sample was pretreated before AN-TPD. An empty-tube test was done prior to AN-TPD to check for possible AN formation. The test result showed no N2O formation during temperature ramping, indicating that the system was sufficiently insulated to prevent any significant AN formation during the NH3+NO2 feed to the SGB. Meanwhile, the N2 (m/e = 28) signal evolved while 200 ppm of NO2/NH3 was fed, suggesting that NO2–SCR occurred during ionization within MS.

2.4.4. In Situ DRIFTS Measurements

In situ DRIFTS was performed to monitor surface nitrite/nitrate and NH3 interaction with the prepared samples (i.e., degreened and SO2- and SO3-exposed samples). A VERTEX 70 FTIR spectrometer (Bruker) equipped with a liquid N2-cooled mercury cadmium telluride (MCT) detector was used to acquire the IR spectra. Degreened powder samples or crushed SOx-exposed monoliths were loaded into a sample cup inside a reaction cell. A KBr bed was set on the bottom of the sample cup, and then the remaining volume of the sample cup was filled with sample. The reaction cell was covered well and mounted with a Ca2F window. A thermocouple was installed at the center of a sample bed to monitor the sample bed temperature during the measurement. The temperature of the sample bed was controlled at 150 °C with a PID regulator (Eurotherm 2416), and MFCs (Bronkhorst Hi-Tech) were set to supply the required gas mixture to the reaction cell. 1% H2O + Ar was fed as base feed throughout the measurement. The required gases were fed in the following steps: (1) heating to 150 °C in base feed only and (2) 400 ppm of NO2 + base feed for 1.5 h. Meanwhile, DRIFTS spectra were measured (accumulation of 60 scans, resolution of 4 cm–1) over time under 100 N mL·min–1 of total flow. The outlet gas leaving the reaction cell was continuously monitored with mass spectrometry (Hidden HPR-20 QUI MS). Note that the reported temperature in the DRIFTS test was monitored by a thermocouple located at the center of the sample bed. However, the heating plate located under the sample cup was set at 242 °C. Thus, a temperature gradient resulting from thermal radiation was present and expected.

2.4.5. First-Principles Calculations

Spin-polarized density functional theory (DFT) calculations were performed using the Vienna Ab initio Simulation Package (VASP).61,62 The Kohn–Sham orbitals were expanded with plane waves with a 480 eV cutoff value, and the interaction between the core and valence electrons was described by the projector augmented wave (PAW)63,64 method. The valence electrons treated for each atom were Cu(11), Si(4), S(6), Al(3), O(6), and H(1). The k-point sampling was restricted to the gamma point. The Perdew–Burke–Ernzerhof (PBE)65 functional was used to describe the exchange-correlation effects. To account for van der Waals interactions, the Grimme 3D approach66 was applied and a Hubbard U-term was used to describe the localized Cu 3d electrons, with the U value set to 6 eV.

The convergence criteria for the self-consistent field (SCF) loop was set to 10–5 eV, and structures were considered to be at a minimum if the norm of all forces was less than 0.02 eV/Å. The transition states were located using the Climbing Image Nudged Elastic Band (CI-NEB) approach67,68 and was confirmed by a single imaginary frequency. Vibrational analysis was performed using the finite difference method. To sample low energy configurations, ab initio molecular dynamics (AIMD) simulations were carried out at 300 K using a Nosé–Hoover thermostat in the NVT ensemble.69,70 Several structures along the trajectories were extracted and relaxed to determine the low-energy structures. Bader charge analysis was performed using the implementation from the Henkelman group.71,72 To describe the CHA cage, a hexagonal unit cell containing 36 Si and 72 O was used with the experimental lattice constants (13.8026 Å, 13.8026 Å, 15.0753 Å). For each Cu ion introduced into the unit cell, one Si atom was replaced by one Al atom for charge compensation.

3. Results

3.1. First-Principles Calculation on SOx Adsorption over Cu Sites

Two experimentally observed resting states of the Cu2+ ions in CHA are Z2Cu and ZCuOH.26,73 Depending on the sample composition and pretreatment, other Cu2+ sites may form and it is important to consider which Cu sites may be present under the conditions investigated. In our study, the samples have been exposed to SOx at 400 °C in the presence of O2 and H2O. At 400 °C, Cu2+ is mainly bound to the framework and we would therefore expect to have both Z2Cu and ZCuOH present. In addition, the presence of O2 could lead to the formation of Cu dimers as copper ions can possibly migrate, forming a pair of Cu2+ ions, which can adsorb O2 forming Z2CuO2Cu complexes.74 Other similar dimer sites have been proposed such as Z2Cu–O–Cu and Z2CuO2H2Cu.30,75,76 The Cu sites that will be investigated in the computational study are Z2Cu, ZCuOH, and Z2CuO2Cu.

To explore the effect of SO2 and SO3, we calculated reaction pathways for the formation of CuHSO3 and CuHSO4 on Z2Cu, ZCuOH, and Z2CuO2Cu. The reaction of SO2 on Z2Cu and ZCuOH has previously been studied using DFT calculations by Jangjou et al.,45 and they found that the addition of SO2 results in the formation of more stable CuHSO3 complexes on ZCuOH than on Z2Cu. In addition, they found that the formation of CuHSO3 exhibited a lower activation barrier when it was formed on ZCuOH.45 Furthermore, the stability of CuHSOx complexes from the reaction of SO2 and SO3 with ZCuOH has been considered with DFT calculations13 and it was found that adsorption of SO3 gives rise to more stable complexes than SO2 adsorption. However, the understanding of the reaction of SO3 with ZCuOH and Z2Cu and the role of Z2CuO2Cu upon sulfur exposure is currently limited. Herein, we investigate the reaction pathway for the combinations of SO2 and SO3 on Z2Cu, ZCuOH, and Z2CuO2Cu, which are Cu sites expected to be present during the first sulfation step in our study.

The reaction of SO2 and SO3 with Z2Cu is calculated in the presence of a single adsorbed H2O molecule. If there is no adsorbed H2O molecule on Z2Cu, it was not possible to form any stable structures with SO2 and SO3, which is in agreement with a previous observation by Jangjou et al.45 for SO2 reacting with Z2Cu. For the reaction of SO2 with Z2Cu (Figure 1a), H2O adsorbs on the Cu ion with a Cu–O bond length of 1.95 Å and an adsorption energy of 0.87 eV (I → II). The introduction of SO2 in the zeolite cage is preferred by 0.55 eV, compared to SO2 in the gas phase; however, we do not find any bond to the Cu ion (II → III). Instead, the oxygen in SO2 coordinates to Cu with a bond length of 2.49 Å. The first step in the reaction of SO2 with Z2Cu–H2O results in the cleavage of the H–OH bond, forming a CuHSO3 complex and a Brønsted acid site (BAS), which is endothermic by 0.52 eV (III → IV). From a Bader charge analysis, sulfur in CuHSO3 is in a +4 oxidation state, as is the case for the gas-phase SO2. In IV, sulfur binds through an OH group, which is not the preferred configuration. In structure IV, the −OH group in CuHSO3 coordinates with the BAS with an SO-HBAS bond length of 3.73 Å. In the next step (IV → V), the CuHSO3 complex is rotated. The barrier for the rotation (0.56 eV) is related to the change in which group is coordinating with the Brønsted site. In structure V, an oxygen from CuHSO3 is coordinating with the Brønsted site resulting in a slight stabilization. The last step is the breakage of the SOH–Cu bond and formation of a SO–Cu bond (V → VI), which is 0.29 eV lower in energy. The final structure is 0.15 eV higher in energy compared to the case with SO2 physisorbed in the cage (structure III). Moreover, as the formation of structure Cu is associated with a high activation barrier, we conclude that the reaction of SO2 with Z2Cu is not favorable.

Figure 1.

Figure 1

Energy landscapes for SO2 and SO3 adsorption onto Z2Cu (a, b) and ZCuOH (c, d) and the subsequent formation of CuHSOx complexes. Atomic color codes: H (white), N (blue), O (red), Al (pink), Si (dark yellow sticks), S (yellow), and Cu (bronze).

SO3 can also react with Z2Cu, as shown in Figure 1b. The first step is also, in this case, the adsorption of H2O (I → II). Introducing SO3 into the cage does not result in adsorption; instead, SO3 is physisorbed (exothermic by 0.51 eV) with one oxygen coordinated to Cu with a bond length of 3.3 Å (II→ III). SO3 can react with H2O, splitting the O–H bond in H2O and forming a CuHSO4 complex and a Brønsted site (III → IV). From Bader charge analysis, sulfur remains in a +6 oxidation state, as is the case for the gas-phase SO3. In structure IV, the OH group in CuHSO4 is coordinating with HBAS. The next step (IV → V) is related to a change of coordination, resulting in the oxygen of CuHSO4 coordinating to HBAS with a bond length of 1.64 Å. The scission of the SOH–Cu bond has a small barrier of 0.11 eV and is preferred by 0.15 eV (V → VI). The last step is the rotation of CuHSO4, resulting in the formation of a second SO–Cu bond (VI → VII). This step has a barrier of 0.41 eV and is exothermic by 0.11 eV. The final structure (VIII) is 0.75 eV more stable compared to the case of having physisorbed SO3 (structure III). When comparing the reaction of SO2 and SO3 with Z2Cu, it becomes clear that SO3 forms ZCuHSO4 complexes that are more stable (0.86 eV in difference) than ZCuHSO3 complexes formed by the reaction with SO2. The highest activation barrier for formation of ZCuHSO4 is 0.41 eV (VI → VII), while the effective barrier is 1.46 eV (III → V) for formation of ZCuHSO3. This suggests that SO3 may result in more severe deactivation than SO2. The higher stability of sulfur species in Cu-CHA with the oxidation state of +6 is in agreement with previous DFT studies.77

The reaction of SOx with ZCuOH can proceed in the absence of H2O. The introduction of SO2 into the CHA cage is preferred by 0.25 eV (I → II, Figure 1c). Next, SO2 can react with ZCuOH forming a CuHSO3 complex that is preferred by 0.14 eV and associated with a small barrier of 0.2 eV (II → III). The last step is the rotation of the sulfur complex, resulting in the breakage of a SOH–Cu bond and, at the same time, the formation of a SO–Cu bond. This step has a barrier of 0.39 eV and is 0.41 eV lower in energy. The stable CuHSO3 complex (structure IV) is 0.55 eV more stable compared to physisorbed SO2 (structure II).

The reaction of SO3 with ZCuOH is shown in Figure 1d. Physisorption of SO3 is exothermic by 0.56 eV (I → II). SO3 can react with ZCuOH forming a CuHSO4 complex that is 0.77 eV more stable and associated with a barrier of 0.16 eV (II → III). The following steps are connected to the reorientation of the CuHSO4 complex away from the eight-membered ring. The first step has a barrier of 0.13 eV and is endothermic by 0.08 eV (III → IV). The second step has a barrier of 0.12 eV and is exothermic by 0.33 eV (IV → V). The last step is the rotation of the CuHSO4 complex, resulting in the breakage of a SOH–Cu bond and the formation of a SO–Cu bond (V → VI). The final CuHSO4 complex (structure VI) is 1.3 eV more stable than physisorbed SO3 (structure II).

For ZCuOH, the same trend as for Z2Cu is observed, where SO3 forms more stable sulfur complexes (−1.3 versus −0.55 eV), which is associated with lower activation barriers compared to the SO2-derived sulfur complexes (0.22 versus 0.39 eV).

As the SOx exposure is performed in the presence of O2 at 400 °C, a pair of Cu ions may adsorb O2 forming framework-bound peroxo species, Z2CuO2Cu.74 The reaction landscape for the reaction of SO2 and SO3 with Z2CuO2Cu is shown in Figure 2. The starting structure I could originate from a pair of Cu+ ions that adsorb O2.74 SO2 cannot adsorb directly onto the complex (structure II). Instead, SO2 can react (low barrier of 0.16 eV) with the adsorbed O2 molecule forming a stable Z2Cu(SO4)Cu complex (II → III), which is 2.68 eV more stable. During the reaction, sulfur is oxidized from +4 to +6. The strongly exothermic nature of sulfur oxidizing for Cu-CHA has been observed previously for SO2 oxidation over the mobile peroxo complex [Cu2O2(NH3)4]2+.57 The Cu ions remain in a +2 oxidation state, supplying the two electrons needed to form the sulfate species (SO2–4). This complex cannot react directly with SO2 or SO3; however, in the presence of H2O, it is possible to form CuHSO3 and CuHSO4 complexes. H2O can adsorb onto Cu with a binding energy of 0.46 eV (III → IV). The next step is the breakage of a SO–Cu bond (IV → V). Furthermore, a bond between Cu and an oxygen in the framework (Ofw) is broken and a new Cu–Ofw bond is formed. The configuration of structure V is favored for the formation of sulfur complexes. The next step (V → VI) is endothermic by 0.26 eV and is the result of the coordination of hydrogen from H2O, which changes from one oxygen to another that is bonded to sulfur. The next steps follow two different paths, which involves the reaction with either SO2 or SO3 indicated in black or red with a star (*), respectively. SO2 (black) can react with H2O, splitting the O–H bond of H2O and forming a CuHSO3 and CuHSO4 complex (VII → VIII), which is endothermic with 0.35 eV. The next step (VIII → IX) is the change of the coordination of an OH group bound to sulfur, which is turned away. The last step is the rotation of CuHSO3 that leads to the breakage of a SOH–Cu bond and formation of a SO–Cu bond (IX → X). This step has a barrier of 0.56 eV and is exothermic by 0.21 eV. From structure VI, it is also possible to follow the structure denoted with an *, which is the reaction with SO3. Here, SO3 can react with H2O forming two CuHSO4 complexes (VII* → VIII*). The formation of CuHSO4 is associated with a small barrier of 0.11 eV and is exothermic by 0.38 eV. As in the previous case, the next step involves the rotation of HSO4, so the −SOH group is no longer bonded to Cu (VIII* → IX*). The rotation has a barrier of 0.4 eV and is exothermic by 0.4 eV.

Figure 2.

Figure 2

Energy landscape for the reaction of SO3 and SO2 with Z2CuO2Cu and the subsequent formation of ZCuHSOx complexes. Atomic color codes are as in Figure 1.

From the reaction landscapes, we conclude that more stable sulfur complexes are formed when reacting Z2CuO2Cu with SO3 compared to SO2. This is similar to the reactions on Z2Cu and ZCuOH. In addition, the activation barriers associated with the reaction with SO3 are lower than for SO2. Compared to Z2Cu and ZCuOH, the first reaction with SO2 (II → III) forms more stable complexes suggesting that Z2CuO2Cu causes more severe deactivation. To form Z2CuO2Cu, the catalyst must be exposed to O2. When SO2 is cofed with O2, severe deactivation has been experimentally observed,49 which is consistent with our calculations. Furthermore, it is possible to oxidize SO2 to SO2–4 over the Z2CuO2Cu and the presence of sulfur in +6 oxidation states has been observed using X-ray absorption spectroscopy.78 The sulfur-derived complexes formed when reacting SOx with Z2CuO2Cu are the same as those for the cases with Z2Cu and ZCuOH. This indicates that even though more sulfur complexes may form on Z2CuO2Cu, it is difficult to unravel the origin of the species. However, the proximity of two CuHSO4 ions could potentially stabilize the complexes. The observation that SO3 forms more stable sulfur complexes that are associated with lower barriers compared to SO2 for all Cu sites is consistent with the observation of more SOx release in TPD experiments.9 In addition, it is observed that SO3 in general causes more deactivation of the low-temperature NH3–SCR.9

3.2. Elemental Analysis, N2 Physisorption, and XRD Analysis of Powder Cu/SSZ-13

Elemental analysis and N2 physisorption were conducted for the prepared copper-exchanged SSZ-13 powders, and the results for the degreened samples are shown in Table S1. SAR (Si/Al molar ratio) was around 14, and the Cu content was 0.83 wt %. The BET surface area decreased (from 711 to 651 m2·g–1) after impregnation of Cu ions. With respect to sulfated samples, 33 and 159 μmol·g–1washcoat of sulfur content were obtained for SO2- and SO3-exposed samples, respectively. In Figure S3, a comparison of X-ray diffractograms for the degreened H and Cu forms of SSZ-13 shows that the CHA framework structure was well maintained without noticeable diffraction peaks of CuxOy crystalline particles (36.30 and 38.72° 2θ)5 or Cu(OH)2 precipitates (12.83 and 25.84° 2θ)37 in the sample powders. Note that the fresh sample was degreened in standard SCR conditions at 750 °C for 5 h and a low Cu loading amount was set (ca. 1 wt % Cu) to avoid possible formation of copper-containing particles. This suggests that no or minor amounts of CuxOy particles were formed and if any was formed; they were below the detection limit of the XRD. It is, thus, assumed that most copper is present as isolated Cu ions interacting with 1 or 2 Alf over the CHA structure of the degreened powder.

The effect of SO2 and SO3 exposure to Cu-CHA was examined using SEM and XRD with the following sulfation sequence: (i) SO2/SO3+O2+H2O; (ii) NH3 + O2+H2O and (iii) SO2/SO3+O2+H2O. We confirmed that no clear changes in crystal morphology from SO2- and SO3-exposed samples using SEM (see Figure S4) were observed compared to the degreened sample. XRD data are shown in Figure 3, and for the degreened sample, catalyst powder was used while for the SO2 and SO3-exposed samples, washcoat was scraped off from the monoliths. In addition, XRD of the cordierite is shown and it is clear that the cordierite present in the samples that were scraped off the monoliths can be neglected. The XRD results show shifted diffraction peaks (Figure 3). The inset (dashed purple) in Figure 3 represents the region of crystallographic planes containing [1 0 0] (ca. 9.5°), [−1 1 0] (ca. 13°), and [1 1 0] (14°).46 Both SO2- and SO3-exposed samples caused peak shifting toward either lower or higher angles compared to the degreened sample, respectively. This suggests that SO2 and SO3 exposure at 400 °C results in changes of the CHA unit cell size but no damaged framework structure according to Bragg’s law.79 Decreased peak intensity for SO2- and SO3-exposed samples indicates that sulfur species were successfully introduced into the CHA unit cell. However, questions remain as to why SO2 and SO3 caused different trends in the shift of the peak toward lower and higher angles, respectively. This different peak shifting trend implies that different lattice distances resulted from SO2 and SO3 exposure with NH3 oxidation as the middle step.

Figure 3.

Figure 3

X-ray diffractogram for degreened and SO2- and SO3-exposed samples. The SO2- and SO3- exposed samples were scraped off the washcoat from monoliths. In addition, XRD from pure cordierite is shown. The inset illustrates X-ray diffraction peak shifting in the position of crystallographic planes with indices of [1 0 0] (ca. 9.5°), [−1 1 0] (ca. 13°), and [1 1 0] (ca. 14°).

UV–visible diffuse reflectance spectroscopy (UV–vis-DRS) was performed to investigate possible formation of different sulfur species. Reference UV spectra were measured for copper compounds (i.e., CuSO4, Cu(OH)2, CuO) and various ammonium (bi)sulfite/sulfate in hydrated and ambient conditions, as shown in Figure S5. Figure 4 presents the resulting UV spectra as a function of wavelength for degreened and SO2- and SO3-exposed samples. Figure 4a shows for all samples two typical UV absorption features corresponding to the d–d transition of hydrated Cu2+ ions (ca. at 800 nm) and the ligand to metal charge transfer (LMCT) transition caused by the isolated Cu2+ ions (lattice O → Cu2+, at ca. 220 nm) of Cu/SSZ-13, as reported in literature.49,80,81 Moreover, UV absorption was observed for both SO2- and SO3-exposed samples over the entire measured wavelength, which was not the case for the degreened sample. This suggests the presence of ammonium (bi)sulfite/sulfate species since these species display considerable UV absorbance over the entire wavelength, as shown in Figure S5b. Both SO2- and SO3-exposed samples show in addition a broader UV absorption band at 250–300 nm compared to the degreened sample. The first derivative of Figure 4a was applied to qualitatively examine the effect of SO2 and SO3 with NH3 oxidation as a middle step, as shown in Figure 4b. The result shows different features of the SO2- and SO3-exposed samples compared to the degreened sample, especially at 220–400 nm. Dissimilar UV absorption features between the SO2- and SO3-exposed samples as in Figure 4c indicates the possible formation of different types of sulfur species interacting with NH3.

Figure 4.

Figure 4

UV–visible diffuse reflectance spectra for degreened and SO2- and SO3-exposed powder samples in terms of (a) absorbance, (b) first derivative of absorbance, and (c) first derivative of absorbance with the degreened sample subtracted.

In comparison with reference bulk materials in Figure S5, CuSO4 (Figure S5a) and various ammonium (bi)sulfites/sulfates (Figure S5b) were considered as potentially formed sulfur species arising during SO2 and SO3 exposure. In the region of 220–400 nm, the broad UV absorbance for CuSO4 and ammonium (bi)sulfite/sulfate indicates that copper sulfate can be potential sulfur species interacting with Cu2+ ions, and the observed additional absorbance (shoulder at 250–300 nm in Figure 4a) is likely to originate from ammonium sulfate ((NH4)2SO4) or ammonium sulfite ((NH4)2SO3). To exclude the background resulting from Cu-CHA with respect to the SO2- and SO3-exposed samples, the UV spectrum of the degreened sample was subtracted from the SO2- and SO3-exposed samples. In Figure 4c, the resulting first derivative of UV absorbance shows the comparison of the derived sulfur species during the SO2 and SO3 exposure with NH3 saturation. By comparing the reference samples (i.e., bulk ammonium (bi)sulfite/sulfate) in Figure S5d, the result implies that the SO2 and SO3 exposures, with NH3 oxidation as the middle step, are associated with the formation of ammonium sulfite ((NH4)2SO3), ammonium sulfate ((NH4)2SO4), or (bi)sulfate ((NH4)HSO4).

3.3. H2 TPR in Degreened and SO2- and SO3-Exposed Cu/SSZ-13

XRD and UV–vis-DRS results suggest that different sulfated species are interacting with NH3 and copper ions inside the CHA unit cell, resulting from SO2 and SO3 exposure of Cu-CHA (with ammonia oxidation as a middle step). This implies that the formed sulfated species can influence the reduction properties of isolated Cu2+ ions. Bergman et al.60 showed that SO2 exposure influences the reduction property of Cu2+ in Cu-CHA using in situ XAS. Thus, H2 temperature-programmed reduction (H2-TPR) was carried out to investigate the reduction property of the SO2- and SO3-exposed samples.

Figure 5a shows H2 consumption as a function of temperature for degreened and SO2- and SO3-exposed samples. The degreened sample shows H2 consumption in the temperature range of 200–650 °C, resulting from Cu2+ species reduction by H2. In general, ZCuOH and Z2Cu are regarded as primarily isolated Cu2+ species over Cu-CHA and the observed H2 consumption peaks are commonly assigned to ZCuOH and Z2Cu at low and high temperatures, respectively, due to different binding strengths of Cu2+ species interacting with the CHA framework.5,48,73,82,83 In addition, in our recent study, combining ab initio calculations with experimental results, we also found the importance Z2CuOCu, Z2CuHOOHCu, and Z2CuOOCu to describe the reduction process.84

Figure 5.

Figure 5

H2 temperature-programmed reduction for degreened and SO2- and SO3-exposed samples showing H2 consumption (a), desorption of SO2 (b), and NH3 release (c), and molar ratios of consumed H2 and desorbed SO2 per Cu content of samples (hydrated condition, gas feed: 0.2% H2/Ar, heating rate: 10 °C·min–1, total flow rate: 20 N mL·min–1).

For the degreened sample in Figure 5a, the main reduction peak at 430 °C was impeded due to sulfur species, implying an interaction between sulfur and copper species. In contrast to the degreened sample, the sulfated samples show larger H2 consumption. SO2 and NH3 signals were detected as shown in Figure 5b,c, suggesting that SO2 and NH3 originated from decomposition of the remaining ammonium (bi)sulfite/sulfate species. The SO2-exposed sample shows two distinct reduction features during H2 TPR at ca. 293 and 652 °C (Figure 5a). For the SO3-exposed sample, relatively intense reduction peaks were observed at 657 °C with a shoulder (604 °C) and at 293 °C. The SO3-exposed sample showed three distinct H2 consumption peaks at 293, 604, and 657 °C, but only two reduction peaks corresponding to ca. 293 and 652 °C were observed for the SO2-exposed sample. Thus, it is likely that the peak at 293 °C resulted from hydrogen reacting with ammonium sulfite/sulfate species and the peaks at 604 and 657 °C originated from sulfur species interacting with Cu2+ ions since SO2 started to be observed above 400 °C with a peak at 577 °C in Figure 5b. In addition, an H2S (mass 34) was measured and a weak signal was detected above 650 °C for the SO3-exposed sample in Figure S6. Based on this, we suggest that the main hydrogen consumption peak at 657 °C might be associated with H2S formation since it requires a two-electron transfer to form H2S as well as cleavage of S–O/S=O bonds to become the sulfide (S2–) of H2S. It should be noted that the detection of H2S in the MS is rather difficult, and we suggest that this is the reason for the significantly smaller and noisier H2S peak compared to the hydrogen consumption peak. In contrast. no H2S signal was observed for the SO2-exposed sample. This is probably because of the lower amount of H2S formation as a result of the relatively lower sulfur content in the SO2-exposed sample compared to the SO3-exposed sample according to the ICP results in Section S3, SI. Both SO2- and SO3-exposed samples showed an SO2 desorption, with a single peak in the same temperature region (peak at 577 °C) (see Figure 5b). However, the sulfur release from the SO3-exposed sample was significantly larger. Note that H2SO4 and SO3 signals were not detected during H2-TPR but these compounds would be challenging to detect in the calorimeter setup if they are in low concentrations. In addition, ammonia desorption was clear from both SO2- and SO3-exposed samples, with two desorption peaks Figure 5c.

The consumed H2 per Cu and SO2 release per Cu from the H2 TPR are shown in Figure 5d. For the degreened sample without sulfur added, the H2/Cu ratio is 0.55, indicating that a one-step reduction of copper occurred. This is common at these temperatures, since a full reduction to metallic copper usually requires higher temperatures. Larger H2 consumption is found for the SO2-exposed sample, which is significantly increased for the SO3-exposed sample, simultaneously with the SO2 desorption increasing. Larger H2/Cu and SO2/Cu ratios are expected if the SOx-exposed sample involves a larger sulfur content. Figure 5d suggests that SO3 exposure results in a much larger sulfur uptake compared to SO2 exposure. This is in line with the ICP analysis results regarding the sulfur content (Section S3, SI). Considering the observed NH3 release and SO2 desorption features in Figure 5b,c, it can be concluded that NH3-sulfur interacting species remained following the sulfation and ammonia exposure step at 400 °C. Different sulfur speciation is expected to result from the SO2 and SO3 exposure, with NH3 as a middle step, forming ammonium sulfite ((NH4)2SO3) and ammonium sulfate ((NH4)2SO4) or (bi)sulfate ((NH4)HSO4) species according to the UV–vis DRS results.

3.4. AN-TPD with Degreened and SO2- and SO3-Exposed Cu/SSZ-13

AN can easily be formed from NO2 and NH3, and during its decomposition, it is well known that N2O is formed. Normally, a larger amount of N2O is reported as the NO2 concentration is increased during SCR reaction.35 We have reported increased N2O formation for Cu-CHA compared to H-CHA, indicating that AN interacts with copper species.5 It is, therefore, expected that SO2 and SO3 influence the AN and N2O formation.

Temperature-programmed desorption with AN (AN-TPD) was carried out for degreened and SO2- and SO3-exposed samples to investigate the effects of SO2 and SO3 on AN and N2O formation. The mass 28 (N2) signal was detected, while 200 ppm of NH3/NO2 was fed for 90 min in the empty-tube test (see Figure S7). This suggests that the NO2 SCR occurred during the ionization process in the mass spectrometer. In addition, no N2O signal was seen during heating when measuring with the FTIR as shown in Figure S7, suggesting that no AN was formed in the reactor system. The results for AN-TPD over Cu/SSZ-13 are presented in Figure 6. The NH3 concentrations during heating are presented in Figure 6a, where the NH3 release was increased for SO2- and SO3-exposed samples compared to the degreened samples. An increased NH3 release is expected due to residual sulfur species, since ammonium (bi)sulfite/sulfate species could form during the NH3+NO2 feed period of the AN-TPD test. These results are in line with the studies in the literature,43,58,85,86 where ammonia was suggested to interact with sulfate species resulting in larger ammonia storage. Notably, the SO3-exposed sample shows the largest NH3 release; furthermore, its NH3 release peak is also shifted from lower to higher temperatures compared to that for the SO2-exposed and degreened samples.

Figure 6.

Figure 6

NH3 (a), NO (b), NO2 (c), and N2O (d) concentrations during ammonium nitrate temperature-programmed desorption for degreened and SO2- and SO3-exposed samples (gas feed: 200 ppm of NH3/NO2 + 5% H2O + Ar at 150 °C during adsorption, heating rate: 10 °C·min–1, total flow rate: 1200 N mL·min–1). Note that the NO concentration is during the adsorption phase.

Figure 6b shows the NO formation curves during NH3+NO2 adsorption for degreened and SO2- and SO3- exposed samples. It has been reported that NO formation originates from NO2 disproportionation and subsequent surface nitrite oxidation by NO2, leading to surface nitrate formation with gas-phase NO release.8789 The ratios for NO formed divided by removed NO2 during the adsorption phase were 0.03, 0.03, and 0.01 for the degreened, SO2-exposed, and SO3-exposed samples, respectively. A ratio of 0.33 is expected due to the disproportionation reaction where nitrates are formed.89 Thus, these values are significantly lower and the reason for this is likely that we also have slow NO2 SCR occurring at this temperature, since we use a quite high temperature (150 °C) and a large amount of the NO2 that is consumed is due to the SCR reaction. NO formation was decreased for both SO2-exposed and, in particular, SO3-exposed samples compared to the degreened sample. This suggests that surface nitrate storage was decreased due to sulfur species interacting with copper species since sulfur species do not show noticeable interaction with Brønsted sites.45 In this respect, it is postulated that most surface nitrate was formed on copper species under these conditions (presence of water, for example), since significantly reduced NO formation was observed for the SO3-exposed sample containing the largest sulfur content. Indeed, lower NO2 formation during the heating is found (Figure 6c), which is in line with the decreased surface nitrate storage, as NO2 formation is expected from the decomposition of the formed surface nitrate species during the NH3+NO2 adsorption.

Figure 6d shows N2O formation during heating for the degreened and SO2- and SO3-exposed samples. N2O formation due to thermal AN decomposition is commonly observed.90 The N2O curve from the SO2-exposed sample shifted toward slightly lower temperatures compared to the degreened sample. Moreover, a similar amount of N2O was released from the SO2-exposed sample (31.9 μmol·g–1washcoat) as for the degreened sample (31.3 μmol·g–1washcoat). It is slightly more N2O amount for the SO2-exposed sample (↑ 1.9%), but the differences could be within the accuracy of the experiments. Despite similar N2O formation, the NO2 release was clearly decreased for the SO2-exposed sample (Figure 6c) as compared to that of the degreened sample, which we interpret as a result of nitrate decomposition. Thus, these results overall indicate that AN formation also decreased for the SO2-exposed sample. However, it should be noted that some of the released NO2 could react in the slow NO2 SCR reaction. In contrast to the SO2-exposed sample, the SO3-exposed sample showed a remarkably decreased N2O release (14.6 μmol·g–1washcoat) corresponding to a ca. 53% reduction in N2O release compared to the degreened sample. The N2O peak was also shifted toward lower temperatures compared to the SO2-exposed and degreened samples. This suggests that SO3 exposure has a stronger influence on N2O formation compared to SO2 exposure.

It should be mentioned that sulfur species such as SO2, SO3, H2SO4, and H2S were measured by FTIR during AN-TPD. Among the sulfur species, SO2 and H2SO4 were observed but the SO3 and H2S were below the detection limit. However, a slightly negative SO2 signal (below 0 ppm) was observed from the SO2- and SO3-exposed samples that was symmetrical with the positive H2SO4 curve, as shown in Figure S8. Thus, it is likely that there were overlapping bands in the FTIR that were slightly overcompensated by the software for SO2 and H2SO4 according to Figure S9. To ease interpretation, the H2SO4 and SO2 signals were merged as SO2+H2SO4 (blue curve in Figure S8), and sulfur species are denoted in this paragraph. The desorption of these sulfur species features was still maintained, showing three desorption peaks at 250, 368, and 537 °C for the SO2-exposed sample and a monotonically increasing SO2+H2SO4 curve with a peak at 533 °C for the SO3-exposed sample. The marked different desorption features for these sulfur species suggests that SO3 exposure causes more stable sulfur species compared to SO2 exposure of Cu-CHA. For the SO2-exposed sample, the H2SO4 peaks at 250 and 368 °C were located at same positions with two NH3 release peaks at 180–215 min, as shown in Figure 6a. On the other hand, the SO3-exposed sample showed an even larger NH3 release; however, there was no significant desorption of sulfur species during NH3 release as compared to the SO2-exposed sample.

3.5. In Situ DRIFTS Measurement with Degreened and SO2- and SO3-Exposed Cu-CHA

AN-TPD results showed decreased NO formation during the NH3+NO2 feed for the SO2- and SO3-exposed samples, implying that surface nitrite oxidation with NO2 was severely limited. Especially, the SO3-exposed sample led to a considerably lower N2O formation. However, the SO2-exposed sample gave comparable N2O formation compared to the degreened sample, although less NO2 formation was observed, consistent with that less AN was formed, as evident from the NO formation in the adsorption phase. To corroborate on these results, surface nitrite/nitrate species were investigated with DRIFTS in the presence of 400 ppm of NO2 + 1% H2O + Ar at 150 °C. It should be mentioned that a relatively higher concentration of H2O is usually introduced to mimic the real exhaust environment. However, 1% H2O was set as the base feed here along with the background feed (Ar) since this was the maximum water feed in our DRIFTS setup. In addition, note that the powder sample was used for the degreened case, while crushed monoliths were used for the SOx-exposed monoliths. Thus, the sulfur exposed samples contain a large amount of cordierite and the spectra were therefore normalized with the amount of copper they contained.

Figure 7 shows the NO2 adsorption phase for the degreened AND SO2- and SO3-exposed samples at 150 °C. The resulting spectral evolutions are illustrated in the region 2000–1000 cm–1 to investigate the surface adsorption species. The surface nitrate region (1800–1500 cm–1) is magnified in the right panel from the original spectrum marked with the dashed line box. The addition of 400 ppm of the NO2 + base feed mixture (Figure 7a) results in peaks related to the formation of chelating bidentate ZCu2+-(NO3) structures (1750–1500 cm–1).91 The four bands at 1624, 1608, 1597, and 1575 cm–1 are associated with υ(NO) vibrations of different types of chelating bidentate nitrates.91,92 The peak at 1608 cm–1 reached the highest intensity, when NO2 was saturated (orange curve). Typical NO+ formation (∼2155 cm–1)93 was not observed here under an excess of water, due to H2O interaction with Brønsted sites Si(OH)Al. The negative band (∼1453 cm–1) is associated with H2O removal.94 The band 1898 cm–1 was assigned as NO interaction with Cu2+ (i.e., Cu2+-NO).95 Note that the negative peak (∼1453 cm–1) may affect the neighboring nitrate feature.

Figure 7.

Figure 7

Spectrum evolution during NO2 adsorption in DRIFTS for the (a) degreened, (b) SO2-exposed, and (c) SO3-exposed samples with (d) comparison of the NO2 spectra of panels (a), (b), and (c) after 2 min of NO2+H2O exposure. Spectral evolution with gray represents intermediate spectra prior to NO2 saturation (orange). Spectra in all panels were normalized by the number of moles of copper within the sample cup (gas feed: 400 ppm of NO2 + 1% H2O + Ar at 150 °C, total flow rate: 100 N mL·min–1, sample bed temp.: 150 °C).

For the SO2-exposed sample in Figure 7b, surface nitrate features were observed over the region 2000–1000 cm–1. The four bands associated with nitrate species (1629, 1611, 1597, and 1576 cm–1) were developed within 5 min and then consumed. Moreover, bands 1887 (positive band) and 1441 cm–1 (negative band) were observed as well as the degreened sample. Surprisingly, at 10 min (blue curve), a decrease in the surface nitrate species peaks was observed, which were further decreased until NO2 saturation (orange curve). During NO2 adsorption, a decrease in the nitrates is not expected. The negative band at 1441 cm–1, due to water removal,94 is quite large. It is possible that there is an overlap between the 1441 cm–1 negative band and the nitrate bands; thus, when the water removal increases, it decreases the bands in the nitrate region. This is not observed for the degreened sample, and the reason for that could be that the nitrate peaks are significantly larger; thus, the interference with the band at 1441 cm–1 might be less. For the SO3-exposed sample in Figure 7c, a similar trend was observed as the SO2-exposed sample. For example, the surface nitrate region increased at 5 min (green curve), followed by decreased intensities until NO2 saturation (orange curve). Likewise, for the degreened sample, the positive band at 1885 and negative band at 1439 cm–1 were observed. In Figure 7d, a comparison is done between the samples after only 2 min of NO2+H2O exposure, where the water peaks for the SO2- and SO3-exposed samples are still quite small. From these results, it is clear that the nitrate bands are decreased for SO2- and SO3-exposed samples compared to the degreened sample.

To summarize, the DRIFT results show that there are fewer nitrates on the SO2- and SO3-exposed samples, as would be expected from the flow reactor experiments. However, it should be mentioned that there could be an overlap with the band for water removal, which could decrease the peak for nitrates.

4. Discussion

At low temperature, surface nitrate formation is observed in the presence of either NO+O2 or NO2 in Cu-CHA.89,91,92,96 An in situ DRIFTS measurement shows that NO2 kinetically promotes surface nitrate formation in Cu-CHA compared to NO+O2 presence in dry conditions.89 In wet conditions, the nitrate species formation is significantly hindered due to water interactions.89 NO release is typically seen during isothermal NO2 adsorption in Cu-CHA, as shown in Figure 6b as well as in the literature,87,88,97,98 resulting from surface nitrite oxidation by NO2, and thereby, NO is formed through the following reactions (R.1 and R.2) below. According to Ruggeri et al.,89 when H2O is present, Cu dimers are likely dissociated into two ZCuOH via R.3.

4. R.1
4. R.2
4. R.3

Ruggeri et al.89 derived R.1R.3 by NO or NO2 exposure to Cu-CHA at 120 or 150 °C in dry or wet conditions using in situ DRIFTS. Cu dimers in R.1 are critically important as a surface oxygen provider, initiating surface nitrite formation and leading to the additional nitrate in the form of copper nitrate and NO via R.2. A recent study performed by Negri et al.96 reported that possible structures of nitrate species and their local environment show that only ZCuOH in the 8-MR within the CHA cage is the active copper species forming the framework-interacting chelating bidentate nitrate (i.e., Z[Cu2+(NO3)]).96 Other nitrate candidates such as bridging bidentate or monodentate nitrate configurations were difficult to reconcile the experimental features under exposure of Cu-CHA to an NO+O2 mixture.91,96

The SO2 and SO3 exposure to Cu-CHA significantly influences the surface nitrate formation. In the AN TPD (Figure 6b), the NO release during isothermal NH3+NO2 adsorption is significantly decreased. Since NO production during NO2 exposure is associated with nitrate formation, according to reactions R.1 and R.2,89 these results suggest that the surface nitrate formation is hindered due to resulting sulfur species after SO2 and SO3 treatment of Cu-CHA. The lowest NO release was observed for the SO3-exposed sample (see Figure 6b), suggesting that this sample exhibited the lowest amount of nitrates, which is consistent with the fact that it contained the largest amount of sulfur, as seen in the ICP measurements (Section S3, SI). These results are also consistent with the in situ DRIFT experiments (see Figure 7d). Our first-principles calculations shown in Figure 1b,d describe that the SO3 interaction with both Z2Cu and ZCuOH is possible. These findings are in agreement with ICP analysis showing an S/Cu molar ratio of 1.2, which indicates that all copper sites were occupied by sulfur species. In contrast, SO2 primarily interacts with ZCuOH but not Z2Cu as suggested by first-principles calculations (Figure 1a,c). Since ICP results show low sulfur storage for the SO2-exposed sample, it suggests that there is a low amount of ZCuOH in our sample. These results are consistent with our H2 TPR (Figure 5a) where the lower reduction peak, usually assigned to ZCuOH, is small for the degreened sample. A deconvolution of hydrogen TPR suggests only 15% ZCuOH. This value is slightly lower than the theoretical value according to Paolucci et al.,26 which is 35%. The reason for this could be that in our synthesis of SSZ-13, we use Na cations and, according to Lee et al.,99 SSZ-13 synthesis with the Na cation can increase 2Al sites over CHA. In our study, SSZ-13 was synthesized using Na and TMAda+, and this could result in a larger amount of 2Al, which increases Z2Cu and thereby decreases ZCuOH.

In terms of speciation of sulfur according to the first-principles calculation in Figures 1 and 2, SO2 interaction with ZCuOH results in (bi)sulfite (ZCuHSO3) and SO2 interactions with dicopper cause both (bi)sulfite and (bi)sulfate surface species. Meanwhile, the interaction of SO3 with Z2Cu and ZCuOH leads to only (bi)sulfate (ZCuHSO4). The (bi)sulfate resulting from the SO3 exposure shows a lower energy compared to the other sulfur species (i.e., ZCuHSO3) derived from the SO2 exposure. Importantly, SO2 and SO3 interaction with Cu dimers shows a further decrease in energy compared to the SO2 and SO3 adsorption with monomeric copper species (i.e., Z2Cu and ZCuOH). The difference between the interaction of SO2 and SO3 with Cu dimers is that SO2 adsorption causes both ZCuHSO3 and ZCuHSO4. The SO3 adsorption, however, results in only two ZCuHSO4. In this respect, two H2 consumption peaks at 604 and 657 °C can be associated with (bi)sulfate interacting with copper monomers and dimers, respectively, in Figure 5a.

As mentioned above and described in reaction R.3, in the presence of water, ZCuOH is the active species for the surface nitrate formation in the form of copper nitrate (ZCuNO3) by following reactions below (R.4R.6), as described in the refs (89,100). In the feed of only NO2, NO+ on the copper site results from NO2 disproportionation with forming nitric acid (HNO3) via R.4. In R.6, the resulting copper-nitrite oxidation by NO2 leads to copper nitrate and NO. For the SO3-exposed sample, most copper sites are occupied by (bi)sulfate in the form of copper-(bi)sulfate (ZCuHSO4). ZCuHSO4, therefore, deactivates the surface nitrate formation by blocking the copper site from the access of the NO+ or HNO3, resulting in the lowest NO release during the adsorption phase in AN-TPD. In the same manner, decreased NO release was seen for the SO2-exposed sample compared with the degreened sample. However, more NO release was observed due to less sulfur content compared to the SO3-exposed sample.

4. R.4
4. R.5
4. R.6

In AN-TPD, stored surface nitrate species are typically desorbed in the form of NO2 for the degreened sample, as can be seen in Figure 6c. For the SO2- and SO3-exposed samples, decreased NO2 release was seen because of decreased nitrate storage resulting from sulfur species, especially the SO3-exposed sample. With respect to N2O, the SO3-exposed sample released the lowest amount of N2O (14.6 μmol·g–1washcoat) compared to the SO2-exposed sample (31.9 μmol·g–1washcoat).

Surprisingly, a comparable amount of N2O was released for the SO2-exposed sample as degreened sample (31.3 μmol·g–1washcoat), although decreased nitrate storage was clearly observed via NO formation in Figure 6b. This suggests that sulfur may increase N2O selectivity during NO2 SCR. Sulfur enhances ammonia storage on the catalyst. We hypothesize that the increased availability of ammonia on the catalyst for the SO2 poisoned sample could result in that more N2O is formed compared to NO2 release. According to first-principle calculations performed by Feng et al.,36 the decomposition of the reaction intermediate (H2NNO) over the Cu site leads to N2O and H2O. Their suggested specific copper-configuration is NH3-solvated Cu dimers. In Figure S8, the observed SO2+H2SO4 release and N2O production (Figure 6d) show that for the SO2 poisoned sample, there is sulfur release already from 200 °C. The main N2O desorption occurs at 300 °C, suggesting that some of the copper sites were regenerated for the SO2-exposed sample; thus, resulting H2NNO intermediates can possibly lead to N2O through the regenerated copper sites. This combination with the increased availability of ammonia could be one possible explanation. However, for the SO3-exposed sample, there was significantly more sulfur stored on the sample as seen in ICP experiments. In addition, the sulfate species were very stable, which is seen by the fact that the SO2+H2SO4 release occurred at significantly higher temperature and was not detected during N2O formation. These results suggest that more copper sites were occupied by sulfur species when N2O was formed and in addition significantly less nitrates were formed in the adsorption phase, which could explain the low N2O formation for the SO3-exposed sample.

5. Conclusions

NO2+NH3 titration to form AN and subsequent temperature-programmed desorption was applied for the degreened and SO2- and SO3-exposed Cu-CHA to investigate the effect of SO2 and SO3 on surface nitrate and N2O formation. Sulfation was carried out at 400 °C to mimic the scenario of SO2/SO3 contamination of the SCR catalyst resulting from SO2 oxidation to SO3/H2SO4 over the DOC or coated filter. H2-TPR and in situ DRIFTS were applied to investigate the reduction properties and surface species response in the presence of sulfur species derived by SO2- and SO3-exposure treatments. Critical effects of SO2 and SO3 on Cu-CHA, its surface nitrate speciation, and N2O formation were found. The following major findings were established:

  • 1.

    First-principles calculations demonstrate that SO2 adsorption on monocopper sites results in (bi)sulfite and dicopper causes both (bi)sulfite and (bi)sulfate surface species. SO3 adsorption on mono- and dicopper results in only (bi)sulfate. Sulfur species derived from the SO3 exposure are lower in energy compared to sulfur species derived from the SO2 exposure. In addition, SOx adsorption on the dicopper exhibits higher thermodynamic stability as compared to SOx adsorption on the monocopper in Cu-CHA, indicating that speciation of copper has a significant effect on the speciation and thermal stability of sulfur species over Cu-CHA.

  • 2.

    SO3 exposure causes a more detrimental contamination of Cu-CHA compared to SO2 exposure, according to the ICP and H2-TPR results, in line with an earlier reported SO2/SO3-exposure study9 as well as showing good agreement with our theoretical calculations. The SO2 exposure mainly shows sulfur contamination with ZCuOH, but the SO3 exposure results in the sulfation not only with ZCuOH but also with Z2Cu. Therefore, the SO3-exposed Cu-CHA shows a larger sulfur uptake compared to the SO2-exposed Cu-CHA.

  • 3.

    In situ DRIFTS demonstrated that similar surface species result from reactions between NO2 and sulfur species derived from the SO2 and SO3 exposures. After 2 min of NO2 exposure, significantly lower bands were observed in the nitrate region for the SO2- and SO3-exposed samples compared to the degreened sample. The results from DRIFT could explain why NO release was decreased for both SO2- and SO3-exposed samples during NH3+NO2 adsorption in AN-TPD. Especially, the SO3-exposed sample showed the lowest NO formation, suggesting that surface nitrate formation was significantly hindered due to (bi)sulfate occupying the copper sites in the form of copper-(bi)sulfate (ZCuHSO4).

  • 4.

    AN-TPD showed similar N2O release for the SO2-exposed Cu-CHA compared to the degreened Cu-CHA, while a lower NO2 release for SO2-exposed Cu-CHA. These results suggest that SO2 exposure can increase the N2O selectivity compared to the degree end Cu-CHA during AN decomposition. In contrast, SO3-exposed Cu-CHA shows significantly lower N2O formation compared to the degreened and SO2-exposed Cu-CHA, due to its highly stable sulfur species (i.e., Cu-(bi)sulfate). SO3 exposure caused more stable sulfur species compared to SO2 exposure. The stability and amount of the sulfur species are likely important factors that influence N2O and N2 selectivity since SO2-exposed and SO3-exposed samples shows totally different N2O formation properties.

Acknowledgments

The Competence Centre for Catalysis is hosted by Chalmers University of Technology and financially supported by the Swedish Energy Agency (Project No. 52689-1) and the member companies Johnson Matthey, Perstorp, Powercell, Preem, Scania CV, Umicore, and Volvo Group. JDB and HG acknowledge support from the European Union’s Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie grant agreement No. 955839 (CHASS). The calculations have been performed at PDC (Stockholm) and NSC (Linköping) through a SNIC grant.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsengineeringau.4c00004.

  • Hydrothermal synthesis of Cu-CHA and washcoating on the monolith substrate, SO2 and SO3 exposure to the Cu-CHA procedure with SO3 calibration, elemental analysis results from ICP, N2 physisorption and XRD, SEM images of zeolite crystal morphology, UV–vis DRS measurement for bulk sulfur-containing compounds, H2S detection during H2-TPR, DRIFTS, empty-tube test during the adsorption step, and SO2/H2SO4/N2 release during heating in AN-TPD (PDF)

Author Contributions

CRediT: Joonsoo Han conceptualization, data curation, formal analysis, investigation, methodology, writing-original draft; Joachim Dithmer Bjerregaard conceptualization, data curation, formal analysis, investigation, writing-original draft; Henrik Grönbeck conceptualization, funding acquisition, methodology, supervision, writing-review & editing; Derek Creaser conceptualization, methodology, supervision, writing-review & editing; Louise Olsson conceptualization, funding acquisition, methodology, supervision, writing-review & editing.

The authors declare no competing financial interest.

Supplementary Material

References

  1. Chen H.-Y.; Wei Z.; Kollar M.; Gao F.; Wang Y.; Szanyi J.; Peden C. H. F. A comparative study of N2O formation during the selective catalytic reduction of NOx with NH3 on zeolite supported Cu catalysts. J. Catal. 2015, 329, 490–498. 10.1016/j.jcat.2015.06.016. [DOI] [Google Scholar]
  2. Gao F.; Walter E. D.; Kollar M.; Wang Y.; Szanyi J.; Peden C. H. F. Understanding ammonia selective catalytic reduction kinetics over Cu/SSZ-13 from motion of the Cu ions. J. Catal. 2014, 319, 1–14. 10.1016/j.jcat.2014.08.010. [DOI] [Google Scholar]
  3. Gao F.; Kwak J. H.; Szanyi J.; Peden C. H. F. Current Understanding of Cu-Exchanged chabazite Molecular Sieves for Use as Commercial Diesel Engine deNOx Catalysts. Top. Catal. 2013, 56 (15), 1441–1459. 10.1007/s11244-013-0145-8. [DOI] [Google Scholar]
  4. Shih A. J.; González J. M.; Khurana I.; Ramírez L. P.; Peña L A.; Kumar A.; Villa A. L. Influence of ZCuOH, Z2Cu, and Extraframework CuxOy Species in Cu-SSZ-13 on N2O Formation during the Selective Catalytic Reduction of NOx with NH3. ACS Catal. 2021, 11 (16), 10362–10376. 10.1021/acscatal.1c01871. [DOI] [Google Scholar]
  5. Han J.; Wang A.; Isapour G.; Härelind H.; Skoglundh M.; Creaser D.; Olsson L. N2O Formation during NH3-SCR over Different Zeolite Frameworks: Effect of Framework Structure, Copper Species, and Water. Ind. Eng. Chem. Res. 2021, 60 (49), 17826–17839. 10.1021/acs.iecr.1c02732. [DOI] [Google Scholar]
  6. Biomass—renewable energy from plants and animals. U.S. Energy Information Administration. [Google Scholar]
  7. Szymaszek A.; Samojeden B.; Motak M. The Deactivation of Industrial SCR Catalysts—A Short Review. Energies 2020, 13 (15), 3870. 10.3390/en13153870. [DOI] [Google Scholar]
  8. Dahlin S.; Englund J.; Malm H.; Feigel M.; Westerberg B.; Regali F.; Skoglundh M.; Pettersson L. J. Effect of biofuel- and lube oil-originated sulfur and phosphorus on the performance of Cu-SSZ-13 and V2O5-WO3/TiO2 SCR catalysts. Catal. Today 2021, 360, 326–339. 10.1016/j.cattod.2020.02.018. [DOI] [Google Scholar]
  9. Hammershøi P. S.; Jangjou Y.; Epling W. S.; Jensen A. D.; Janssens T. V. W. Reversible and irreversible deactivation of Cu-CHA NH3-SCRcatalysts by SO2 and SO3. Appl. Catal. B: Environ. 2018, 226, 38–45. 10.1016/j.apcatb.2017.12.018. [DOI] [Google Scholar]
  10. Xi Y.; Ottinger N.; Su C.; Liu Z. G. Sulfur Poisoning of a Cu-SSZ-13 SCR Catalyst under Simulated Diesel Engine Operating Conditions. SAE Int. J. Adv. & Curr. Prac. in Mobility 2021, 3 (5), 2690–2694. 10.4271/2021-01-0576. [DOI] [Google Scholar]
  11. Mesilov V. V.; Bergman S. L.; Dahlin S.; Xiao Y.; Xi S.; Zhirui M.; Xu L.; Chen W.; Pettersson L. J.; Bernasek S. L. Differences in oxidation-reduction kinetics and mobility of Cu species in fresh and SO2-poisoned Cu-SSZ-13 catalysts. Applied Catalysis B: Environmental 2021, 284, 119756 10.1016/j.apcatb.2020.119756. [DOI] [Google Scholar]
  12. Mesilov V.; Dahlin S.; Bergman S. L.; Xi S.; Han J.; Olsson L.; Pettersson L. J.; Bernasek S. L. Regeneration of sulfur-poisoned Cu-SSZ-13 catalysts: Copper speciation and catalytic performance evaluation. Applied Catalysis B: Environmental 2021, 299, 120626 10.1016/j.apcatb.2021.120626. [DOI] [Google Scholar]
  13. Hammershøi P. S.; Vennestrøm P. N. R.; Falsig H.; Jensen A. D.; Janssens T. V. W. Importance of the Cu oxidation state for the SO2-poisoning of a Cu-SAPO-34 catalyst in the NH3-SCR reaction. Appl. Catal. B: Environ. 2018, 236, 377–383. 10.1016/j.apcatb.2018.05.038. [DOI] [Google Scholar]
  14. Hammershøi P. S.; Jensen A. D.; Janssens T. V. W. Impact of SO2-poisoning over the lifetime of a Cu-CHA catalyst for NH3-SCR. Appl. Catal. B: Environ. 2018, 238, 104–110. 10.1016/j.apcatb.2018.06.039. [DOI] [Google Scholar]
  15. Auvray X.; Arvanitidou M.; Högström Å.; Jansson J.; Fouladvand S.; Olsson L. Comparative Study of SO2 and SO2/SO3 Poisoning and Regeneration of Cu/BEA and Cu/SSZ-13 for NH3 SCR. Emission Control Science and Technology 2021, 7 (4), 232–246. 10.1007/s40825-021-00203-4. [DOI] [Google Scholar]
  16. Chen Y.-R.; Wei L.; Kumar A.; Wang D.; Epling W. S. How changes in Cu-SSZ-13 catalytic oxidation activity via mild hydrothermal aging influence sulfur poisoning extents. Catalysis Science & Technology 2022, 12 (22), 6891–6902. 10.1039/D2CY01394K. [DOI] [Google Scholar]
  17. Cheng Y.; Montreuil C.; Cavataio G.; Lambert C. The Effects of SO2 and SO3 Poisoning on Cu/Zeolite SCR Catalysts. SAE J. 2009, 10.4271/2009-01-0898. [DOI] [Google Scholar]
  18. Majewski W. A.Diesel Oxidation Catalyst; DieselNet Technology Guide: DieselNet. [Google Scholar]
  19. Kumar A.; Li J.; Luo J.; Joshi S.; Yezerets A.; Kamasamudram K.; Schmidt N.; Pandya K.; Kale P.; Mathuraiveeran T.. Catalyst Sulfur Poisoning and Recovery Behaviors: Key for Designing Advanced Emission Control Systems; SAE Technical Paper Series, 2017.
  20. Granestrand J.; Dahlin S.; Immele O.; Schmalhorst L.; Lantto C.; Nilsson M.; París R. S.; Regali F.; Pettersson L. J.. Catalytic aftertreatment systems for trucks fueled by biofuels – aspects on the impact of fuel quality on catalyst deactivation. In Catalysis : Vol.30, Vol. 30; The Royal Society of Chemistry, 2018; pp 64–145. [Google Scholar]
  21. Majewski W. A.Diesel Filter Regeneration, date of access: 2024–05–13. DieselNet Technology Guide: DieselNet.. [Google Scholar]
  22. Kumar A.; Li J.; Luo J.; Joshi S.; Yezerets A.; Kamasamudram K.; Schmidt N.; Pandya K.; Kale P.; Mathuraiveeran T. Catalyst Sulfur Poisoning and Recovery Behaviors: Key for Designing Advanced Emission Control Systems. SAE Int. 2017, 10.4271/2017-26-0133. [DOI] [Google Scholar]
  23. Wu Y.; Andana T.; Wang Y.; Chen Y.; Walter E. D.; Engelhard M. H.; Rappé K. G.; Wang Y.; Gao F.; Menon U.; Daya R.; Trandal D.; An H.; Zha Y.; Kamasamudram K.; et al. A comparative study between real-world and laboratory accelerated aging of Cu/SSZ-13 SCR catalysts. Appl. Catal. B Environ. 2022, 318, 121807 10.1016/j.apcatb.2022.121807. [DOI] [Google Scholar]
  24. Paolucci C.; Khurana I.; Parekh A. A.; Li S.; Shih A. J.; Li H.; Di Iorio J. R.; Albarracin-Caballero J. D.; Yezerets A.; Miller J. T.; et al. Dynamic multinuclear sites formed by mobilized copper ions in NO x selective catalytic reduction. Science 2017, 357 (6354), 898–903. 10.1126/science.aan5630. [DOI] [PubMed] [Google Scholar]
  25. Liu C.; Malta G.; Kubota H.; Kon K.; Toyao T.; Maeno Z.; Shimizu K.-i. In Situ/Operando IR and Theoretical Studies on the Mechanism of NH3–SCR of NO/NO2 over H–CHA Zeolites. J. Phys. Chem. C 2021, 125 (25), 13889–13899. 10.1021/acs.jpcc.1c03431. [DOI] [Google Scholar]
  26. Paolucci C.; Parekh A. A.; Khurana I.; Di Iorio J. R.; Li H.; Albarracin Caballero J. D.; Shih A. J.; Anggara T.; Delgass W. N.; Miller J. T.; et al. Catalysis in a Cage: Condition-Dependent Speciation and Dynamics of Exchanged Cu Cations in SSZ-13 Zeolites. J. Am. Chem. Soc. 2016, 138 (18), 6028–6048. 10.1021/jacs.6b02651. [DOI] [PubMed] [Google Scholar]
  27. Göltl F.; Bhandari S.; Lebrón-Rodríguez E. A.; Gold J. I.; Zones S. I.; Hermans I.; Dumesic J. A.; Mavrikakis M. Identifying hydroxylated copper dimers in SSZ-13 via UV-vis-NIR spectroscopy. Catalysis Science & Technology 2022, 12 (9), 2744–2748. 10.1039/D2CY00353H. [DOI] [Google Scholar]
  28. Paolucci C.; Di Iorio J. R.; Schneider W. F.; Gounder R. Solvation and Mobilization of Copper Active Sites in Zeolites by Ammonia: Consequences for the Catalytic Reduction of Nitrogen Oxides. Acc. Chem. Res. 2020, 53 (9), 1881–1892. 10.1021/acs.accounts.0c00328. [DOI] [PubMed] [Google Scholar]
  29. Li H.; Paolucci C.; Khurana I.; Wilcox L. N.; Goltl F.; Albarracin-Caballero J. D.; Shih A. J.; Ribeiro F. H.; Gounder R.; Schneider W. F. Consequences of exchange-site heterogeneity and dynamics on the UV-visible spectrum of Cu-exchanged SSZ-13. Chem. Sci. 2019, 10 (8), 2373–2384. 10.1039/C8SC05056B. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Ipek B.; Wulfers M. J.; Kim H.; Göltl F.; Hermans I.; Smith J. P.; Booksh K. S.; Brown C. M.; Lobo R. F. Formation of Cu2O2 2+ and Cu2O 2+ toward C–H Bond Activation in Cu-SSZ-13 and Cu-SSZ-39. ACS Catal. 2017, 7 (7), 4291–4303. 10.1021/acscatal.6b03005. [DOI] [Google Scholar]
  31. Wang G.; Zhi C.; Wang Y.; Wang Q. Theoretical Investigation on the Structure-Activity Relationship for Methane Activation in Cu-CHA Zeolite. Comput. Theor. Chem. 2023, 114228 10.1016/j.comptc.2023.114228. [DOI] [Google Scholar]
  32. Khurana I.; Albarracin-Caballero J. D.; Shih A. J. Identification and quantification of multinuclear Cu active sites derived from monomeric Cu moieties for dry NO oxidation over Cu-SSZ-13. J. Catal. 2022, 413, 1111–1122. 10.1016/j.jcat.2022.08.005. [DOI] [Google Scholar]
  33. Gao F.; Mei D.; Wang Y.; Szanyi J.; Peden C. H. Selective Catalytic Reduction over Cu/SSZ-13: Linking Homo- and Heterogeneous Catalysis. J. Am. Chem. Soc. 2017, 139 (13), 4935–4942. 10.1021/jacs.7b01128. [DOI] [PubMed] [Google Scholar]
  34. Liu B.; Yao D.; Wu F.; Wei L.; Li X.; Wang X. Experimental Investigation on N2O Formation during the Selective Catalytic Reduction of NOx with NH3 over Cu-SSZ-13. Ind. Eng. Chem. Res. 2019, 58 (45), 20516–20527. 10.1021/acs.iecr.9b03294. [DOI] [Google Scholar]
  35. Xi Y.; Ottinger N. A.; Keturakis C. J.; Liu Z. G. Dynamics of low temperature N2O formation under SCR reaction conditions over a Cu-SSZ-13 catalyst. Appl. Catal. B: Environ. 2021, 294, 120245 10.1016/j.apcatb.2021.120245. [DOI] [Google Scholar]
  36. Feng Y.; Janssens T. V. W.; Vennestrøm P. N. R.; Jansson J.; Skoglundh M.; Grönbeck H. The Role of H+- and Cu+-Sites for N2O Formation during NH3-SCR over Cu-CHA. J. Phys. Chem. C 2021, 125 (8), 4595–4601. 10.1021/acs.jpcc.0c11008. [DOI] [Google Scholar]
  37. Shih A. J.; González J. M.; Khurana I.; Ramírez L. P.; Peña L A.; Kumar A.; Villa A. L. Influence of ZCuOH, Z2Cu, and Extraframework CuxOy Species in Cu-SSZ-13 on N2O Formation during the Selective Catalytic Reduction of NOx with NH3. ACS Catal. 2021, 0 (0), 10362–10376. 10.1021/acscatal.1c01871. [DOI] [Google Scholar]
  38. Gao F.; Peden C. H. F. Recent Progress in Atomic-Level Understanding of Cu/SSZ-13 Selective Catalytic Reduction Catalysts. Catalysts 2018, 8 (4), 140. 10.3390/catal8040140. [DOI] [Google Scholar]
  39. Shan Y.; He G.; Du J.; Sun Y.; Liu Z.; Fu Y.; Liu F.; Shi X.; Yu Y.; He H. Strikingly distinctive NH3-SCR behavior over Cu-SSZ-13 in the presence of NO2. Nat. Commun. 2022, 13 (1), 4606. 10.1038/s41467-022-32136-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Kubota H.; Liu C.; Toyao T.; Maeno Z.; Ogura M.; Nakazawa N.; Inagaki S.; Kubota Y.; Shimizu K.-i. Formation and Reactions of NH4NO3 during Transient and Steady-State NH3-SCR of NOx over H-AFX Zeolites: Spectroscopic and Theoretical Studies. ACS Catal. 2020, 10 (3), 2334–2344. 10.1021/acscatal.9b05151. [DOI] [Google Scholar]
  41. Liu C.; Malta G.; Kubota H.; Toyao T.; Maeno Z.; Shimizu K.-i. Mechanism of NH3–Selective Catalytic Reduction (SCR) of NO/NO2 (Fast SCR) over Cu-CHA Zeolites Studied by In Situ/Operando Infrared Spectroscopy and Density Functional Theory. J. Phys. Chem. C 2021, 125 (40), 21975–21987. 10.1021/acs.jpcc.1c06651. [DOI] [Google Scholar]
  42. Chen L.; Janssens T. V. W.; Vennestrøm P. N. R.; Jansson J.; Skoglundh M.; Grönbeck H. A Complete Multisite Reaction Mechanism for Low-Temperature NH3-SCR over Cu-CHA. ACS Catal. 2020, 10 (10), 5646–5656. 10.1021/acscatal.0c00440. [DOI] [Google Scholar]
  43. Kumar A.; Kamasamudram K.; Currier N.; Yezerets A. Effect of Transition Metal Ion Properties on the Catalytic Functions and Sulfation Behavior of Zeolite-Based SCR Catalysts. SAE International Journal of Engines 2017, 10 (4), 1604–1612. 10.4271/2017-01-0939. [DOI] [Google Scholar]
  44. Jangjou Y.; Ali M.; Chang Q.; Wang D.; Li J.; Kumar A.; Epling W. S. Effect of SO2 on NH3 oxidation over a Cu-SAPO-34 SCR catalyst. Catalysis Science & Technology 2016, 6 (8), 2679–2685. 10.1039/C5CY02212F. [DOI] [Google Scholar]
  45. Jangjou Y.; Do Q.; Gu Y.; Lim L.-G.; Sun H.; Wang D.; Kumar A.; Li J.; Grabow L. C.; Epling W. S. Nature of Cu Active Centers in Cu-SSZ-13 and Their Responses to SO2 Exposure. ACS Catal. 2018, 8 (2), 1325–1337. 10.1021/acscatal.7b03095. [DOI] [Google Scholar]
  46. Shih A. J.; Khurana I.; Li H.; González J.; Kumar A.; Paolucci C.; Lardinois T. M.; Jones C. B.; Albarracin Caballero J. D.; Kamasamudram K.; Yezerets A.; Delgass W. N.; Miller J. T.; Villa A. L.; Schneider W. F.; Gounder R.; Ribeiro F. H.; et al. Spectroscopic and kinetic responses of Cu-SSZ-13 to SO2 exposure and implications for NOx selective catalytic reduction. Appl. Catal. A: Gen. 2019, 574, 122–131. 10.1016/j.apcata.2019.01.024. [DOI] [Google Scholar]
  47. Jangjou Y.; Wang D.; Kumar A.; Li J.; Epling W. S. SO2 Poisoning of the NH3-SCR Reaction over Cu-SAPO-34: Effect of Ammonium Sulfate versus Other S-Containing Species. ACS Catal. 2016, 6 (10), 6612–6622. 10.1021/acscatal.6b01656. [DOI] [Google Scholar]
  48. Wang A.; Olsson L.. Insight into the SO2 poisoning mechanism for NOx removal by NH3-SCR over Cu/LTA and Cu/SSZ-13. Chemical Engineering Journal 2020, 395. 125048 10.1016/j.cej.2020.125048. [DOI] [Google Scholar]
  49. Wijayanti K.; Xie K.; Kumar A.; Kamasamudram K.; Olsson L. Effect of gas compositions on SO2 poisoning over Cu/SSZ-13 used for NH3-SCR. Applied Catalysis B: Environmental 2017, 219, 142–154. 10.1016/j.apcatb.2017.07.017. [DOI] [Google Scholar]
  50. Su W.; Li Z.; Zhang Y.; Meng C.; Li J. Identification of sulfate species and their influence on SCR performance of Cu/CHA catalyst. Catalysis Science & Technology 2017, 7 (7), 1523–1528. 10.1039/C7CY00302A. [DOI] [Google Scholar]
  51. Zhang Y.; Zhu H.; Zhang T.; Li J.; Chen J.; Peng Y.; Li J. Revealing the Synergistic Deactivation Mechanism of Hydrothermal Aging and SO2 Poisoning on Cu/SSZ-13 under SCR Condition. Environ. Sci. Technol. 2022, 56 (3), 1917–1926. 10.1021/acs.est.1c06068. [DOI] [PubMed] [Google Scholar]
  52. Shih A. J.; Khurana I.; Li H.; González J.; Kumar A.; Paolucci C.; Lardinois T. M.; Jones C. B.; Albarracin Caballero J. D.; Kamasamudram K.; et al. Spectroscopic and kinetic responses of Cu-SSZ-13 to SO2 exposure and implications for NOx selective catalytic reduction. Applied Catalysis A: General 2019, 574, 122–131. 10.1016/j.apcata.2019.01.024. [DOI] [Google Scholar]
  53. Luo J.; Wang D.; Kumar A.; Li J.; Kamasamudram K.; Currier N.; Yezerets A. Identification of two types of Cu sites in Cu/SSZ-13 and their unique responses to hydrothermal aging and sulfur poisoning. Catal. Today 2016, 267, 3–9. 10.1016/j.cattod.2015.12.002. [DOI] [Google Scholar]
  54. Hammershøi P. S.; Godiksen A. L.; Mossin S.; Vennestrøm P. N. R.; Jensen A. D.; Janssens T. V. W. Site selective adsorption and relocation of SOx in deactivation of Cu–CHA catalysts for NH3-SCR. React. Chem. Eng. 2019, 4 (6), 1081–1089. 10.1039/C8RE00275D. [DOI] [Google Scholar]
  55. Molokova A. Y.; Borfecchia E.; Martini A.; Pankin I. A.; Atzori C.; Mathon O.; Bordiga S.; Wen F.; Vennestrøm P. N. R.; Berlier G.; et al. SO2 Poisoning of Cu-CHA deNOx Catalyst: The Most Vulnerable Cu Species Identified by X-ray Absorption Spectroscopy. JACS Au 2022, 0 (0), 787. 10.1021/jacsau.2c00053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Cheng Y.; Lambert C.; Kim D. H.; Kwak J. H.; Cho S. J.; Peden C. H. F. The different impacts of SO2 and SO3 on Cu/zeolite SCR catalysts. Catal. Today 2010, 151 (3–4), 266–270. 10.1016/j.cattod.2010.01.013. [DOI] [Google Scholar]
  57. Bjerregaard J. D.; Votsmeier M.; Grönbeck H. Mechanism for SO2 poisoning of Cu-CHA during low temperature NH3-SCR. J. Catal. 2023, 417, 497–506. 10.1016/j.jcat.2022.12.023. [DOI] [Google Scholar]
  58. Wijayanti K.; Leistner K.; Chand S.; Kumar A.; Kamasamudram K.; Currier N. W.; Yezerets A.; Olsson L. Deactivation of Cu-SSZ-13 by SO2 exposure under SCR conditions. Catalysis Science & Technology 2016, 6 (8), 2565–2579. 10.1039/C5CY01288K. [DOI] [Google Scholar]
  59. Bergman S. L.; Dahlin S.; Mesilov V. V.; Xiao Y.; Englund J.; Xi S.; Tang C.; Skoglundh M.; Pettersson L. J.; Bernasek S. L. In-situ studies of oxidation/reduction of copper in Cu-CHA SCR catalysts: Comparison of fresh and SO2-poisoned catalysts. Applied Catalysis B: Environmental 2020, 269, 118722 10.1016/j.apcatb.2020.118722. [DOI] [Google Scholar]
  60. Bergman S. L.; Dahlin S.; Mesilov V. V.; Xiao Y.; Englund J.; Xi S.; Tang C.; Skoglundh M.; Pettersson L. J.; Bernasek S. L.. In-situ studies of oxidation/reduction of copper in Cu-CHA SCR catalysts: Comparison of fresh and SO2-poisoned catalysts. Applied Catalysis B: Environmental 2020, 269. 118722 10.1016/j.apcatb.2020.118722. [DOI] [Google Scholar]
  61. Kresse G.; Furthmüller J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B Condens Matter 1996, 54 (16), 11169–11186. 10.1103/PhysRevB.54.11169. [DOI] [PubMed] [Google Scholar]; From NLM
  62. Kresse G.; Hafner J. Ab initio molecular-dynamics simulation of the liquid-metal--amorphous-semiconductor transition in germanium. Phys. Rev. B 1994, 49 (20), 14251–14269. 10.1103/PhysRevB.49.14251. [DOI] [PubMed] [Google Scholar]
  63. Blöchl P. E. Projector augmented-wave method. Phys. Rev. B 1994, 50 (24), 17953–17979. 10.1103/PhysRevB.50.17953. [DOI] [PubMed] [Google Scholar]
  64. Kresse G.; Joubert D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59 (3), 1758–1775. 10.1103/PhysRevB.59.1758. [DOI] [Google Scholar]
  65. Perdew J. P.; Burke K.; Ernzerhof M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77 (18), 3865–3868. 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]; From NLM
  66. Grimme S.; Antony J.; Ehrlich S.; Krieg H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132 (15), 154104 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]; From NLM
  67. Mills G.; Jónsson H.; Schenter G. K. Reversible work transition state theory: application to dissociative adsorption of hydrogen. Surf. Sci. 1995, 324 (2), 305–337. 10.1016/0039-6028(94)00731-4. [DOI] [Google Scholar]
  68. Henkelman G.; Jónsson H. Improved tangent estimate in the nudged elastic band method for finding minimum energy paths and saddle points. J. Chem. Phys. 2000, 113 (22), 9978–9985. 10.1063/1.1323224. [DOI] [Google Scholar]
  69. Nosé S. A unified formulation of the constant temperature molecular dynamics methods. J. Chem. Phys. 1984, 81 (1), 511–519. 10.1063/1.447334. [DOI] [Google Scholar]
  70. Hoover W. G. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A Gen Phys. 1985, 31 (3), 1695–1697. 10.1103/PhysRevA.31.1695. [DOI] [PubMed] [Google Scholar]; From NLM
  71. Tang W.; Sanville E.; Henkelman G. A grid-based Bader analysis algorithm without lattice bias. J. Phys.: Condens. Matter 2009, 21 (8), 084204 10.1088/0953-8984/21/8/084204. [DOI] [PubMed] [Google Scholar]
  72. Yu M.; Trinkle D. R. Accurate and efficient algorithm for Bader charge integration. J. Chem. Phys. 2011, 134 (6), 064111 10.1063/1.3553716. [DOI] [PubMed] [Google Scholar]
  73. Hun Kwak J.; Zhu H.; Lee J. H.; Peden C. H. F.; Szanyi J. Two different cationic positions in Cu-SSZ-13?. Chem. Commun. 2012, 48 (39), 4758–4760. 10.1039/c2cc31184d. [DOI] [PubMed] [Google Scholar]
  74. Wang X.; Chen L.; Vennestrøm P. N. R.; Janssens T. V. W.; Jansson J.; Grönbeck H.; Skoglundh M. Direct measurement of enthalpy and entropy changes in NH3 promoted O2 activation over Cu–CHA at low temperature. ChemCatChem. 2021, n/a (n/a), 2577. 10.1002/cctc.202100253. [DOI] [Google Scholar]
  75. Rhoda H. M.; Plessers D.; Heyer A. J.; Bols M. L.; Schoonheydt R. A.; Sels B. F.; Solomon E. I. Spectroscopic Definition of a Highly Reactive Site in Cu-CHA for Selective Methane Oxidation: Tuning a Mono-μ-Oxo dicopper(II) Active Site for Reactivity. J. Am. Chem. Soc. 2021, 143 (19), 7531–7540. 10.1021/jacs.1c02835. [DOI] [PMC free article] [PubMed] [Google Scholar]; note = PMID: 33970624
  76. Bregante D. T.; Wilcox L. N.; Liu C.; Paolucci C.; Gounder R.; Flaherty D. W. Dioxygen Activation Kinetics over Distinct Cu Site Types in Cu-chabazite Zeolites. ACS Catal. 2021, 11 (19), 11873–11884. 10.1021/acscatal.1c03471. [DOI] [Google Scholar]
  77. Mesilov V.; Xiao Y.; Dahlin S.; Bergman S. L.; Pettersson L. J.; Bernasek S. L. First-Principles Calculations of Condition-Dependent Cu/Fe Speciation in Sulfur-Poisoned Cu- and Fe-SSZ-13 Catalysts. J. Phys. Chem. C 2021, 125 (8), 4632–4645. 10.1021/acs.jpcc.1c01016. [DOI] [Google Scholar]
  78. Mesilov V.; Dahlin S.; Bergman S. L.; Hammershøi P. S.; Xi S.; Pettersson L. J.; Bernasek S. L. Insights into sulfur poisoning and regeneration of Cu-SSZ-13 catalysts: in situ Cu and S K-edge XAS studies. Catal. Sci. Technol. 2021, 11 (16), 5619–5632. 10.1039/D1CY00975C. [DOI] [Google Scholar]
  79. Waseda Y.; Matsubara E.; Shinoda K.. X-Ray Diffraction Crystallography; Springer Berlin: Heidelberg, 2011. 10.1007/978-3-642-16635-8. [DOI] [Google Scholar]
  80. Giordanino F.; Vennestrom P. N.; Lundegaard L. F.; Stappen F. N.; Mossin S.; Beato P.; Bordiga S.; Lamberti C. Characterization of Cu-exchanged SSZ-13: a comparative FTIR, UV-Vis, and EPR study with Cu-ZSM-5 and Cu-beta with similar Si/Al and Cu/Al ratios. Dalton Trans 2013, 42 (35), 12741–12761. 10.1039/c3dt50732g. [DOI] [PubMed] [Google Scholar]
  81. Korhonen S. T.; Fickel D. W.; Lobo R. F.; Weckhuysen B. M.; Beale A. M. Isolated Cu2+ ions: active sites for selective catalytic reduction of NO. Chem. Commun. (Camb) 2011, 47 (2), 800–802. 10.1039/C0CC04218H. [DOI] [PubMed] [Google Scholar]
  82. Gao F.; Washton N. M.; Wang Y.; Kollár M.; Szanyi J.; Peden C. H. F. Effects of Si/Al ratio on Cu/SSZ-13 NH3-SCR catalysts: Implications for the active Cu species and the roles of Brønsted acidity. J. Catal. 2015, 331, 25–38. 10.1016/j.jcat.2015.08.004. [DOI] [Google Scholar]
  83. Gao F.; Kwak J. H.; Szanyi J.; Peden C. H. F. Current Understanding of Cu-Exchanged chabazite Molecular Sieves for Use as Commercial Diesel Engine deNOx Catalysts. Top. Catal. 2013, 56 (15–17), 1441–1459. 10.1007/s11244-013-0145-8. [DOI] [Google Scholar]
  84. Bjerregaard J. D.; Han J.; Creaser D.; Olsson L.; Grönbeck H. On the Interpretation of H2-TPR from Cu-CHA using First-Principles Calculations. J. Phys. Chem. C 2023, 4525. 10.1021/acs.jpcc.3c07998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Wang A.; Olsson L. Insight into the SO2 poisoning mechanism for NOx removal by NH3-SCR over Cu/LTA and Cu/SSZ-13. Chemical Engineering Journal 2020, 395, 125048 10.1016/j.cej.2020.125048. [DOI] [Google Scholar]
  86. Dahlin S.; Lantto C.; Englund J.; Westerberg B.; Regali F.; Skoglundh M.; Pettersson L. J. Chemical aging of Cu-SSZ-13 SCR catalysts for heavy-duty vehicles – Influence of sulfur dioxide. Catal. Today 2019, 320, 72–83. 10.1016/j.cattod.2018.01.035. [DOI] [Google Scholar]
  87. Colombo M.; Nova I.; Tronconi E. A comparative study of the NH3-SCR reactions over a Cu-zeolite and a Fe-zeolite catalyst. Catal. Today 2010, 151 (3–4), 223–230. 10.1016/j.cattod.2010.01.010. [DOI] [Google Scholar]
  88. Villamaina R.; Liu S.; Nova I.; Tronconi E.; Ruggeri M. P.; Collier J.; York A.; Thompsett D. Speciation of Cu Cations in Cu-CHA Catalysts for NH3-SCR: Effects of SiO2/AlO3 Ratio and Cu-Loading Investigated by Transient Response Methods. ACS Catal. 2019, 9 (10), 8916–8927. 10.1021/acscatal.9b02578. [DOI] [Google Scholar]
  89. Ruggeri M. P.; Nova I.; Tronconi E.; Pihl J. A.; Toops T. J.; Partridge W. P. In-situ DRIFTS measurements for the mechanistic study of NO oxidation over a commercial Cu-CHA catalyst. Applied Catalysis B: Environmental 2015, 166–167, 181–192. 10.1016/j.apcatb.2014.10.076. [DOI] [Google Scholar]
  90. Park J.; Lin M. C. Thermal Decomposition of Gaseous Ammonium Nitrate at Low Pressure: Kinetic Modeling of Product Formation and Heterogeneous Decomposition of Nitric Acid. J. Phys. Chem. A 2009, 113 (48), 13556–13561. 10.1021/jp9058005. [DOI] [PubMed] [Google Scholar]; note = PMID: 19845384
  91. Negri C.; Hammershoi P. S.; Janssens T. V. W.; Beato P.; Berlier G.; Bordiga S. Investigating the Low Temperature Formation of Cu(II) -(N,O) Species on Cu-CHA Zeolites for the Selective Catalytic Reduction of NOx. Chemistry 2018, 24 (46), 12044–12053. 10.1002/chem.201802769. [DOI] [PubMed] [Google Scholar]
  92. Negri C.; Borfecchia E.; Cutini M.; Lomachenko K. A.; Janssens T. V. W.; Berlier G.; Bordiga S. Evidence of Mixed-Ligand Complexes in Cu–CHA by Reaction of Cu Nitrates with NO/NH3 at Low Temperature. ChemCatChem. 2019, 11 (16), 3828–3838. 10.1002/cctc.201900590. [DOI] [Google Scholar]
  93. Chen H.-Y.; Wei Z.; Kollar M.; Gao F.; Wang Y.; Szanyi J.; Peden C. H. F. NO oxidation on zeolite supported Cu catalysts: Formation and reactivity of surface nitrates. Catal. Today 2016, 267, 17–27. 10.1016/j.cattod.2015.11.039. [DOI] [Google Scholar]
  94. Bordiga S.; Regli L.; Lamberti C.; Zecchina A.; Bjørgen M.; Lillerud K. P. FTIR Adsorption Studies of H2O and CH3OH in the Isostructural H-SSZ-13 and H-SAPO-34: Formation of H-Bonded Adducts and Protonated Clusters. J. Phys. Chem. B 2005, 109 (16), 7724–7732. 10.1021/jp044324b. [DOI] [PubMed] [Google Scholar]; note = PMID: 16851897
  95. Zhang R.; McEwen J.-S.; Kollár M.; Gao F.; Wang Y.; Szanyi J.; Peden C. H. F. NO Chemisorption on Cu/SSZ-13: A Comparative Study from Infrared Spectroscopy and DFT Calculations. ACS Catal. 2014, 4 (11), 4093–4105. 10.1021/cs500563s. [DOI] [Google Scholar]
  96. Negri C.; Martini A.; Deplano G.; Lomachenko K. A.; Janssens T. V. W.; Borfecchia E.; Berlier G.; Bordiga S. Investigating the role of Cu-oxo species in Cu-nitrate formation over Cu-CHA catalysts. Phys. Chem. Chem. Phys. 2021, 18322. 10.1039/D1CP01754C. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Grossale A.; Nova I.; Tronconi E.; Chatterjee D.; Weibel M. The chemistry of the NO/NO2–NH3 “fast” SCR reaction over Fe-ZSM5 investigated by transient reaction analysis. J. Catal. 2008, 256 (2), 312–322. 10.1016/j.jcat.2008.03.027. [DOI] [Google Scholar]
  98. 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 (1), 243–255. 10.1016/j.cattod.2012.09.002. [DOI] [Google Scholar]
  99. Lee S.; Nimlos C. T.; Kipp E. R.; Wang Y.; Gao X.; Schneider W. F.; Lusardi M.; Vattipalli V.; Prasad S.; Moini A.; et al. Evolution of Framework Al Arrangements in CHA Zeolites during Crystallization in the Presence of Organic and Inorganic Structure-Directing Agents. Cryst. Growth Des. 2022, 22, 6275. 10.1021/acs.cgd.2c00856. [DOI] [Google Scholar]
  100. Cui Y.; Gao F. Cu Loading Dependence of Fast NH3-SCR on Cu/SSZ-13. Emission Control Science and Technology 2019, 5 (2), 124–132. 10.1007/s40825-019-00117-2. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials


Articles from ACS Engineering Au are provided here courtesy of American Chemical Society

RESOURCES