A new popular transition metal-based catalyst: SmMn2O5 mullite-type oxide
English
A new popular transition metal-based catalyst: SmMn2O5 mullite-type oxide
-
1. Introduction
Catalyst plays an important role in the development of human society, from the chemicals production to energy utilization and environmental problem solving [1]. Among them, platinum group metals (PGMs), including Pt, Pd and Rh, are utilized broadly in catalytic fields, arising from their significant catalytic activity, high thermal stability and chemical toxicity resistance [2,3]. The catalysts based on the PGMs are therefore widely used in the fields containing the automobile catalytic converters, pharmaceutical industry, hydrogen production, oil refining, and fine chemical synthesis. However, the high price of the PGMs restricts their further applications [2,4,5]. The ever-increasing industry led to explosion of new catalysts with excellent performance, low price and high stability.
Complex metal oxide (ABmOn) presents stable properties in catalysis reactions reasoning from their skeleton structure connected by metal oxygen polyhedron [6]. As a representative complex metal oxide, perovskite-type oxide presents flexible crystal structures, and therefore provides an opportunity to introduce foreign atoms into the initial structure for stability and catalytic performance enhancement [1,6]. The previous reports confirmed the higher activity of doped Ba0.5Sr0.5Co0.8Fe0.2O3-δ and La0.2Sr0.8CoO3-δ catalysts in oxygen evolution reaction (OER) [7,8] and NOx conversion reactions than that of the most advanced Pt catalysts [9].
Mullite-type oxide SmMn2O5, a typical complex metal oxide, was first reported in 2012 as a diesel engine oxidation catalyst with superior performance in NO conversion [10]. This catalyst with a mixed phase of (SrCeMn5O9.83)(SmMn2O5) exhibits a ~45% increasement over commercial Pt catalyst in NO catalytic oxidation reaction at 300 ℃ arising from the high chemical activity of SmMn2O5. In oxygen reduction reaction (ORR), the overpotential of SmMn2O5 reaches to 0.78 V, higher than that of Pt catalyst (0.7 V), and shows better cycle stability than that of Pt [11]. SmMn2O5 exhibits excellent catalysis performance and provides a new option of highly active, stable, and low-cost heterogeneous catalyst; however, to our knowledge, there is no review article covering the comprehensive information of SmMn2O5 and its applications.
Here, we review the latest advances in the physiochemical properties of mullite SmMn2O5, its significant catalytic performance, and its applications (Fig. 1). We begin with a brief discussion of the crystal structure and key physical properties of SmMn2O5, then discuss its catalytic applications, the catalytic activity sources and reaction mechanisms. We follow with a summary of the recent advances in modification methods that improve the catalytic ability of SmMn2O5. Finally, we present a perspective on further innovations for catalytic applications.
Figure 1
2. Structure and properties
2.1 Crystal structure
SmMn2O5 crystallizes in an orthorhombic three-dimensional crystal structure with a space group Pbam. The lattice parameters are a = 7.43 Å, b = 8.59 Å, c = 5.70 Å with α = β = γ = 90° [12–15]. The structure consists of two main units in structure of MnO6 octahedron and MnO5 square-pyramid (Fig. 2a), where the MnO6 octahedra shares edges via O2 and O3 oxygens to form infinite chains along the c axis while the MnO5 pyramid shares edges via O1 to form a dimer unit Mn2O8. The chains are linked together by MnO5 units in the a-b plane through O3 and O4 oxygens, forming a stable crystal structure [12,13].
Figure 2
In this specifical crystal structure, there are two different Mn atoms: Mn4+ locates at the 4f Wyckoff site in the MnO6 distorted octahedron and the Mn3+ occupies the 4h Wyckoff site in the MnO5 distorted square-pyramid [12,13,16]. Despite the onefold Sm3+, four type oxygen atoms exist in the SmMn2O5 bulk whereat the denoted O1, O2, and O4 atoms connect two Mn ions to form Mn-O-Mn structure while an O3 connects a Mn3+ ion and two Mn4+ ions [6,12,13,16]. The presented double crystal units and the polyvalency of the Mn enable the excellent catalytic activity of SmMn2O5.
The rich coordination environment between Mn and O atoms endows abundant oxygen vacancy in the crystal structure. The density functional theory (DFT) calculation reveals that the low formation energies of O3 and O2 vacancies (EVo = 2.85 eV and 3.86 eV, respectively) enable rich oxygen vacancies appearing along the c-axis direction [6]. The oxygen ion conduction therefore can happen through the hopping of the oxygen from the site of the bridging oxygen Obri in MnO5 to the oxygen vacancy in the octahedral chain (Fig. 2b) [17]. Furthermore, the SmMn2O5 has a small oxygen migration barrier of 0.97–2.81 eV based on its rich oxygen vacancies, revealing the excellent oxygen storage and transmission capability of SmMn2O5 [6]. The polyvalency of Mn atoms and excellent oxygen conductivity of SmMn2O5 lay the foundation for further applications in the catalysis such as gas oxidation, water electrolysis, and ORR.
2.2 Physical and chemical properties
In the past decades, the Mn-based mullite as a dielectric material has been widely investigated in magnetic and ferroelectric properties [18–24]. The excellent electrical and thermodynamic behaviors determine the further application in catalysis.
2.2.1 Ferroelectric and semiconductor properties
RMn2O5 (R = Sm-Lu, Y, or Bi) exhibits ferroelectric property at low temperature (25–40 K) [18]. The intrinsic difference of polar molecule between pristine positively and negatively poled ferroelectric surfaces in the adsorption and reaction processes provide the potential to manipulate catalytic reactions and create ferroelectric chemical sensors [25,26]. However, SmMn2O5 is intrinsically pyroelectric that causes a weak spontaneous polarization at room temperature [25]. The change of dielectric constant and loss tangent with frequency of SmMn2O5 confirm the above property (Fig. 3a) [27].
Figure 3
SmMn2O5 is also a p-type indirect gap semiconductor material with a band gap of ~1.0 eV (Fig. 3b) [28–30]. This p-type characteristics of SmMn2O5 is conducive to the adsorption of oxygen, benefiting for its oxidation reaction. The further temperature dependent impedance spectroscopy confirms the charge carriers hopping conduction in this material (Fig. 3c), which is benefit for the high chemical reactivity of SmMn2O5 [27]. The above rich electrical properties ensure the good activities of SmMn2O5 in chemical reactions.
The thermodynamic property of SmMn2O5 closely relates to temperature and partial pressure of oxygen, and exhibits follow thermal dissociation process [31,32]:
(1) (2) (3) Yankin et al. studied the phase equilibria of SmMn2O5 over the temperature ranging from 973 K to 1123 K while pressure ranging from 10−3 Pa to10−16 Pa, respectively [31]. They found that the thermal decomposition temperature of SmMn2O5 increases with higher oxygen partial pressure (Fig. 3d). The high temperature (point a, 1418 K) of thermal dissociation in air confirms the high thermal stability of SmMn2O5. Li et al. also obtained a similar decomposition temperature of 1373 K for SmMn2O5 by employing the first principles simulations (Fig. 3e), which is in good agreement with the value (1439 K) obtained from thermogravimetric analysis and differential scanning calorimetry measurements [33]. The unchanged phase structure of SmMn2O5 at 1323 K for 8 h further confirmed its high thermal stability (Fig. 3f) [34]. The proved good thermal stability enables the wide reaction temperature of SmMn2O5.
3. Origin of catalytic activity and reaction mechanism
In order to clarify the origin of the catalytic properties, including active centers, electronic structures, and metal-oxygen bond strength, a descriptor is defined to explain the catalytic activity, and a discussion on the fundamental mechanisms of catalytic reactions is further presented.
3.1 Metal-oxygen bond strength
The catalytic activity of SmMn2O5 is mainly derived from Mn-Mn dimeric structure [10]. This distinctive active site is more reactive than single Mn active site based on the cooperative electron transfer between the two Mn atoms in the dimer [30]. Here, the filled eg-orbital and the hybridization between O-2p and Mn-3d determine the high catalytic activity of the SmMn2O5, which is different to precious metals that only the half-filled d-band states determine the activity [35].
Li and coworkers defined a general descriptor for AMn2O5 to clarify the origin the catalytic property in NO oxidation reaction by employing soft X-ray absorption spectroscopy and DFT calculations [36]. They showed that the ground-state high-spin configured Mn4+ (Mnoct) and Mn3+ (Mnpyr) can be regarded as t2g3 and (dxy1, dxz1, dyz1, dz2) respectively; dz2 orbital of Mnpyr closing to Fermi level is occupied by a single electron (Fig. 4a) [29]. Compared with t2g orbital of Mn4+, the dz2 orbital of Mn3+ has strong bonding ability with O-2p orbital. The occupy states of dz2-band therefore determines the interaction between Mn and adsorbed O atom but cannot be a catalytic descriptor because of similar d-band shapes and d-band filling of AMn2O5 with different A atoms [36]. The O-K edge absorption spectra showed that the varied pd hybridization caused by A atom between the Mn-3d and the bulk O-2p orbital exhibits linear relationship with NO conversion efficiency (Fig. 4b), revealing that the pd hybridization is the appropriate catalytic descriptor for AMn2O5 catalysts [36,37]. In detail, the weak hybridization between Mnoct-3d and Obulk-2p leads to the energy lifting of the dz2 orbital of Mnpyr and strongly interaction with O*-2p orbitals, while the strong pd hybridization can produce lower d-band position and weak combination with O* (Fig. 4c) [11,36,38,39].
Figure 4
3.2 Reaction mechanisms
3.2.1 Oxidation reaction mechanisms
There are three typical reaction mechanisms for transition metal oxides (TMOs) in oxidation reaction: Langmuir-Hinshel-wood mechanism (LH) [40,41], Eley-Rideal mechanism (ER) [42,43] and Mars-van Krevelen mechanism (MvK) [44,45]. In the LH mechanism, the TMOs serve as a substrate where the reactants simultaneously adsorb and then react with each other to form new molecules (Fig. 5a), while, in the ER mechanism, only one reactant adsorbs on the substrate and reacts with another reactant appearing in gas phase (Fig. 5b). In the MvK mechanism, the surface oxygen atom of TMOs participates in the reaction. Taking the NO oxidation reaction for example, there are mainly two reaction steps listed in MvK (Fig. 5c): (1) The adsorbed NO on substrate is oxidized to NO2 when combing with a lattice surface oxygen; (2) The reduced surface is re-oxidized to the original state through oxygen adsorption and dissociation [6,37,45–47]. In general, multiple reaction routes combining with different mechanisms can be found in one oxidation reaction [48,49].
Figure 5
Among various oxidation reactions of SmMn2O5, the NO oxidation is the most popular reaction and can be used to explore the reaction mechanism. Based on DFT calculations, Wang et al. [10] proposed that NO oxidation reaction followed ER mechanism via the Mn dimer active sites on the SmMn2O5 (110) surface and displayed the general NO oxidation reaction route as shown in Fig. 5b that: (1) O2 prefers adsorbing on the Mn-Mn sites and decomposes into oxygen atoms; (2) The incident NO gas adsorbs on the O site to form monodentate Mn-nitrite (*NO2), and then transforms into bidentate Mn-nitrate (*NO3) when reacts with the other absorbing oxygen; (3) The *NO3 reacts with an oxygen vacancy on the surface and convert to dissociation NO2. In contrast, Zheng et al. [37] believed that NO oxidation was a MvK reaction process through the cooperation of multiple lattice oxygens (denoted Oα, Oβ and Oγ) on SmMn2O5 (010) surface (Fig. 5d). They believed that: (1) The Oβ serves as a reversible anionic redox center where NO adsorbs on the surface when accepting electrons from the catalyst; (2) O2 molecules fill the Oα vacancy (vOα) to stabilize the critical rotation step to produce much more monodentate nitrites; (3) NO2 desorbs on the surface when giving out electrons; (4) Oα serves as an oxidant to help turning NO into NO2 mediated by Oβ and O2 gas.
Despite the above explanation of NO oxidation on SmMn2O5, there are still some disputes over the real reaction process and mechanism. Chen et al. [6] reported a comprehensive explanation about reaction mechanism of SmMn2O5 by studying the oxygen chemistry in bulk phases as well as on the four low index surfaces-(110), (100), (010) and (001) by the DFT calculations. The results shown that the (001)4+ surface and (010)4+ surface significantly promote the total oxidation rate through the synergistic ER and MvK mechanism (Fig. 5e). Significantly, the rate-limiting step of this reaction is the desorption of NO2 on the surface regardless of the NO oxidation [6,10].
3.2.2 Reduction reaction mechanisms
SmMn2O5 is also mostly used as an efficient catalyst in metal-air batteries based on the ORR. The common ORR mechanism can be divided into two categories: adsorbate evolution mechanism (AEM) [50–54] and lattice oxygen-mediated mechanism (LOM) [53,54]. In AEM, a 4e− reaction process coupling with protons happens on the active metallic sites to achieve an ORR without participation of the lattice oxygen. As shown in Fig. 6a, O2 first adsorbs on the electrode surface that is occupied by *OH, and then form to *OO with releasing a OH−. It is further reduced to *OOH by binding protons in solution. Finally, the O-O bond in the *OOH breaks to form *O and OH−. The *O is further protonated to form *OH in the solution [38]. In LOM, the lattice oxygen directly participates in the reaction and does not involve the formation of *OOH (Fig. 6b). The nature of LOM is mainly related to the covalency of the metal-oxygen (M-O) bond in metal oxides; the stronger covalency of the M-O bonds achieves the higher ability to achieve ORR. Theoretically, the AEM or LOM occurs in a reaction can be adjusted by changing the strength of the M-O bonds that depends on the relative position of the metal d and oxygen p bands in the metal oxides [55]. If the d band of the metal locates at higher energy than that of oxygen p band, the AEM will be prioritized for ORR; conversely, if the energy of oxygen p band is higher than that of the metal d band, electrons will be easier to transfer from the oxygen p band to the metal d band along with the formation of ligand holes, enhancing the redox activity of the lattice oxygen in an ORR with changing from AEM to LOM [55,56].
Figure 6
However, SmMn2O5 with multiple ligand units can hold more variable oxygen environment on the terminated surfaces, causing these two mechanisms may fail to precisely describe the ORR process. Wang et al. [38] reported the third ORR mechanism for SmMn2O5 that named labile oxygen participant adsorbate evolving mechanism (LAM). This mechanism can be described that: proton H firstly adsorbs on labile oxygen to form *OH, then O2 replaces *OH to form adsorbed *OO (Fig. 6c), and then is reduced to intermediate *OOH by binding protons in solution, and finally, the O-O bond in *OOH is broken to form *O and *OH to complete the 4e− process. Based on the DFT calculations, the obtained overpotential of ORR is 0.410 V based on the LAM, which is in good agreement with the experimental result of 0.413 V. In contrast, the overpotential values of ORR based on the AEM and LOM are obviously higher than the experimental results. The above results confirm the rationality of the LAM on the explication of ORR for SmMn2O5 [38].
4. Application in catalysis
The variable-valence element Mn, good oxygen conductivity and excellent thermal stability enhance the catalysis performance. In this section, we provide a brief review on the catalytic applications of SmMn2O5 on both catalytic oxidation reaction and catalytic reduction reaction.
4.1 Automobile catalytic converters
Vehicle exhaust mainly includes nitrogen oxides (NOx) (> 90% NO), carbon monoxide (CO), hydrocarbons (HCs) and particulate matter produced by incomplete combustion (mainly soot), which are harmful to the environment and human health [57–61]. One of the most reliable ways to reduce the toxicity of vehicle exhaust gasses and remove harmful substances is the complete oxidation of exhaust components using catalytic compositions on metal or ceramic block carriers. Comparing to the commercial PGM catalysts (such as Pt/Al2O3), SmMn2O5 displays comparable catalytic activity on automobile exhaust, such as comparable or lower light-off temperature T50 (the temperature at 50% conversion) (Figs. 7a–e, middle) and higher conversion for all five oxidation reactions (Figs. 7a–e, bottom). For example, Wang et al. [10] found that SmMn2O5 achieved a maximum conversion efficiency of 79% compared with 62% for Pt/Al2O3 in NO oxidation (Fig. 7a, bottom). Thampy et al. [62] further confirmed that SmMn2O5 has a maximum NO conversion efficiency of 52% at lower temperature (200 ℃) and Li et al. [36] reported that SmMn2O5 can achieve the maximum NO conversion of 75% at 300 ℃. Similarly, Zhao et al. [33] employed SmMn2O5 in CO oxidation and found that the T50 of SmMn2O5 is 130 ℃ while the highest conversion efficiency of 98% is achieved at 180 ℃, showing comparable performance with Pt/Al2O3 (Fig. 7b, bottom). For the catalysis on soot, SmMn2O5 also shows excellent catalytic activity, especially in NOx-assisted soot combustion [63–66]. Feng et al. [64] showed that both the light-off temperature and activation energy in soot combustion based on SmMn2O5 are greatly accelerated under NO + O2 atmosphere (Fig. 7c, top) comparing to only under O2 (Fig. 7c, bottom), —368 ℃ to 518 ℃ and 65 kJ/mol to 135 kJ/mol, respectively. Moreover, SmMn2O5 catalyst possesses excellent soot oxidation efficiency of 90% at 420 ℃, lower than that of Pt/Al2O3 (480 ℃) [64]. SmMn2O5 also displays better ammonia selective catalytic oxidation performance than 1.2%-Pt/γ-Al2O3 samples when used for NH3 removing from the exhaust gas [67]. The other previous reports further confirmed that SmMn2O5 exists superior activity on VOCs (such as ethanol [30], toluene [68] and benzene [69]) with high hydrothermal stability and durability. For instance, the ethanol complete oxidation temperature is 175 ℃ for SmMn2O5, which is much lower than those for the other catalysts (Fig. 7d, bottom) [30]. SmMn2O5 achieves more than 90% conversion efficiency for methane [70], propane [71,72], propylene [73] and other hydrocarbons at 300–600 ℃ (Fig. 7e). These results confirm that SmMn2O5 is an effective catalyst for automotive exhaust oxidation. The good catalytic activities of SmMn2O5 are attributed to the rich active sites on the surface of SmMn2O5 and the unique Mn-Mn dimer structure.
Figure 7
4.2 OER and supercapacitor
SmMn2O5 is also used in OER and energy storage field due to its abundant lattice oxygen and good oxygen conductivity. Rani synthesized SmMn2O5 hollow long nano-cuboids (HLNCs) and nanorods, and used them for OER in water electrolysis and electrochemical supercapacitor applications [74,75]. The results exhibited that the SmMn2O5 has a high specific capacitance of 256 F/g at 5 mV/s scan rate and high OER activity of 0.37 mA/g at 10 mV/s scan rate with low Tafel slope value of 64 mV/dec (Fig. 8). The internal resistance and the charge transfer resistance were respectively estimated as 13 and 355 Ω from EIS spectra (Fig. 8b), which providing the clear view of fast charge transfer and electronic mobility. Furthermore, the fabricated electrode showed a high stability of 100% retention for 3 h in water oxidization reaction (Fig. 8f) [74]. As shown in Fig. 8c, the specific capacitance calculated from GCD curve is 141 F/g at 1 A/g and acts as a good achievement of fabricated SmMn2O5 HLNCs electrode in supercapacitor [74]. These results indicate that SmMn2O5 has unique oxygen storage, and both good charge transfer capacity and discharge capacity, indicating the significant potential applications in the field of energy storage.
Figure 8
4.3 Potentiometric gas sensor
As a p-type semiconductor, the dz2 state with a single electron occupies the orbitals near the Fermi level, resulting in a lower sensing reaction barrier of SmMn2O5 and promoting the oxidation reaction of the gas molecular on the surface. The above excellent electronic structure therefore induces a rapid response behavior of SmMn2O5 as a gas sensor. Zhu et al. [76] employed SmMn2O5 nanofibers (SNFs), nanotube-in-tubes (NT@NTs), and porous nanotubes (PNTs) as sensors for acetone detection. They found that the SmMn2O5 SNFs sensor exhibited rapid response and recovery times as short as 3 s and 8 s at 300 ℃, respectively (Figs. 9a and b). When used as a sensor for methane detection, Yang et al. [77] found that SmMn2O5 exhibited rapid response and recovery rates of methane with different concentrations at 350–500 ℃. The response and recovery times were significantly shortened to 27 s and 33 s for 400 ppm at 400 ℃ (Fig. 9c). Despite the above less reported applications, SmMn2O5 is also a potential semiconductor material for gas detections directly, and it can be further combined with other semiconductors to form heterojunction structures for rapid detection of more hazardous gasses [78–80].
Figure 9
4.4 Metal-air batteries
Similar to catalytic oxidation, the catalytic reduction reaction also involves the breaking of an oxygen molecule into two O atoms, which dependents on the bonding strength of metal and oxygen in the catalyst. SmMn2O5 with moderated Mn-O* bond has great application potential in catalytic reduction due to the volcano curve relationship between Mn-O* bond strength and catalytic performance [11,36,38,39]. In 2016, Wang et al. [11] first explored the catalytic activity of SmMn2O5 for ORR. They employed SmMn2O5 as an electrocatalyst in Mg-air battery and found that it exhibits similar ORR catalytic activity to Pt/C catalyst in 1 mol/L NaCl neutral solutions but has superior stability than Pt/C (Fig. 10a). In addition, they further investigated the electrocatalytic activity of SmMn2O5 in ORR under alkaline conditions (0.1 mol/L KOH) [39]. The catalyst exhibits an overpotential of 0.413 V and revealed the superior stability with only ~5% decay in activity over 20,000 s, exceeding 15% of Pt/C catalyst (Figs. 10b and c). Moreover, SmMn2O5 was also confirmed excellent electrochemical performance as an electrode of Al-air battery [81] and Zn-air battery [82]. Zn-air batteries constructed with SmMn2O5 nanorods exhibit not only higher peak power density (217 mW/cm2) over commercial Pt/C catalyst (190 mW/cm2) but also excellent cycling stability in long-term charging-discharging test over 170 h (Figs. 10d and e). In general, SmMn2O5 is a promising cathode catalyst and provides new options for advanced energy conversion and storage.
Figure 10
5. Strategies for improving the catalytic performance of SmMn2O5
Despite the significant catalytic activity and stability on various chemical reactions, some strategies, including morphology regulation [71,82–85], metal ion doping [29,86,87], oxygen vacancies introduction [16,28,88,89], and composition with other materials [69,63,73,90–95] have been performed to enhance the properties of low-temperature catalysis, gas adsorption strength and electronic conductivity of SmMn2O5.
5.1 Morphology regulation
The common synthetic methods of mullite oxide are co-precipitation and hydrothermal method. The former method relies heavily on calcination conditions and precipitation pH while the latter is greatly affected by the pH of the solution and the type of precursor. We therefore can obtain larger specific surface area (SSA) and more catalytic active sites (Mn3+ and Mn4+) of SmMn2O5 by modulating the morphology of SmMn2O5 based on different reaction conditions.
By changing the pH of precursor solution, Thampy et al. [83] reported a linear relationship between the SSA of SmMn2O5 and pH while Ma et al. [84] obtained single crystal SmMn2O5 nanorods and nanoparticles. Similarly, by utilizing Mn precursors with different anions, Wang et al. [85] synthetized different SmMn2O5 nanostructures with different morphologies (Fig. 11a), such as granular, flower-like, and mixed structures. In addition, Yang et al. [73] proposed a simple selective dissolution method to modify the surface of SmMn2O5 oxide. They found that part of Sm on the surface can be removed after acid treatment, therefore increasing the exposure of Mn4+ and the oxygen vacancy concentration.
Figure 11
In addition to the above methods, the researchers have developed new preparation processes for different application scenarios and achieved the regulation of morphology. SmMn2O5 with unique interconnected macroporous structure was synthesized by the combustion of ethylene glycol and methanol solutions methods, and was successfully applied in NOx-assisted soot combustion (Fig. 11a) [64]. Three-dimensionally ordered macroporous SmMn2O5 catalyst with large pore sizes was also designed and synthesized successfully by applying a carboxyl-modified colloidal crystal templates method. This synthesized SmMn2O5 nanoparticle exhibits high catalytic activity and anti-sintering performance for soot combustion [65]. Furthermore, leaf-like SmMn2O5 nanosheet with large SSA was applied in the combustion catalysis of C3H8 at 243 ℃ and exhibited an over 50% conversion, which is higher than that of 1 wt% Pd/Al2O3 catalyst [71]. Therefore, we can see that the innovation of preparation method can not only improve the catalytic performance of SmMn2O5 by regulating the morphology, but also improve the production efficiency of materials to a certain extent.
5.2 Metal ions doping
The strategy of ion doping is also an important way to enhance the catalytic activity by regulating the electronic structure of materials [96–99]. In SmMn2O5, replacing Sm atom by other rare earth elements benefits for the catalytic activity, thermal stability and oxygen storage capacity. Feng et al. [86] characterized a series of LaxSm1-xMn2O5-δ (x = 0, 0.1, 0.3, 0.5) and found that the introduction of La atom increased the oxygen adsorption and Mn4+ content, benefiting for the decomposition of surface nitrate/nitrite and the dissociation of NO2. The DFT calculations revealed that the Ba doping destabilizes the nitrite (*NO) species and promotes NO oxidation performance, while the doping of Sr and La enhances the catalytic reactivity of the catalysts in a wide temperature range arising from the stronger ability of O2 (*O2) dissociation [87].
Beside the replacement of Sm atom, substitution of the Mn atoms with TMs can also change the local electronic structure, surface chemical coordination environment and therefore the catalytic activity of SmMn2O5. Liu et al. [92] doped SmMn2O5 by 3d transition metal dopants and found that the catalyst exhibits enhanced low-temperature oxidation activity for CO when doping by Fe. They reasoned this property to the reduced dissociation barrier of O2 inducing by the appearance of Mn-Fe heterodimer at the interface (Fig. 11d).
The mixed-phase mullite oxide by SmMn2O5 and perovskite AMnO3-δ, such as AxSm1-xMn2O5-δ (A = Ca/Sr/Ba), can be seen as another strategy of ions doping. Zhao et al. [29] found that the atomic bonding interfaces between the active phase AMnO3-δ and SmMn2O5 exhibited smaller work functions benefiting for the charge transfer and catalytic activity (Figs. 11b and c).
5.3 Oxygen vacancy introduction
Compared to element doping, the introduction of oxygen vacancy is considered to be a simpler and more effective way to improve the activity of catalyst, which can adjust the valence state of Mn and improve the oxygen conductivity and adsorption activity of O2. The highly dispersed SmMn2O5 nanorods with rich oxygen vacancies were used for water electrolysis [75]; the electrochemical characterization displayed that the nanorods presented a high specific capacitance value of 352 F/g at 10 mV/s and a high current density of 178 mA/g with a lower internal resistance of 1.5 Ω, indicating the good electrical conductivity. Dong et al. [16] and Zhao et al. [89] synthesized a series of amorphous mullite SmMn2O5-δ by laser irradiation. The further ORR of the amorphous mullite SmMn2O5-δ revealed that the increased laser pulse induced higher oxygen vacancy concentration, and therefore induced the variation of ratio of Mn3+ and Mn4+ based on the charge balance principle. The volcano curve formed by ORR catalytic activity and the concentration of oxygen vacancies δ proved that the modest oxygen-defective sample with δ = 0.14 and valence of Mn as 3.36 possessed a superior catalytic activity among the reported family of Mn-based oxides, which was comparable to that of Pt/C (Fig. 11e) [16].
5.4 Composited with other materials
Collaborating with other materials to compensate for the inherent performance shortcomings of a material is a smart and popular strategy. As for SmMn2O5, different composite materials have different emphasis on the performance improvement, such as catalytic oxidation performance at low temperature, selectivity and electronic conductivity (Fig. 12a). The previous reports confirm that the excellent catalytic of the pure SmMn2O5 occurs mainly at medium and high temperature ranging of 300–600 ℃. The strategies of supporting metal nanoparticles on SmMn2O5 or building composite catalysts with SmMn2O5 can improve the low-temperature catalytic activity of SmMn2O5. For example, the platinum sub-nanoclusters were tightly anchored onto SmMn2O5 (Pt/SMO) via atomic layer deposition to form a novel composite catalyst with a bifunctional interfacial structure, resulting in a significantly lowered light-off temperature and high performance on CO oxidation [90,92]. As shown in Fig. 12b, Pt/SMO exhibits excellent low temperature activity for CO oxidation with a T50 at ~86 ℃, which is much lower than that of pure SMO (171 ℃) [90]. The PtIWI/SMO (synthesized by incipient wetness impregnation method) and Pt/Al2O3 exhibited poorer activity than Pt/SMO, implying the key role of the interface structure and inherent active sites in SmMn2O5. The SmMn2O5 modified with Ag and Pd also showed the similar property to that of Pt [73]. Similarly, loading non-precious metal oxides MOx (M = Mn, Fe, Co, Ni, Cu) onto SmMn2O5 mullite (MOx/SMO) is an effective approach to enhance its catalytic performance of CO oxidation [91]. The values of activation energy deducting the contribution of SmMn2O5 for MOx/SMO (M = Cu, Ni, Co, Mn) are 43.5, 67.8, 81 and 74.3 kJ/mol respectively, which are close to those values listed in Fig. 12c, indicating that the activity at low temperature can be mainly attributed to the MOx modification over SmMn2O5 but not the SmMn2O5 [91]. The improvement of the low-temperature activity of SmMn2O5 composited with other catalyst is mainly attributed to the new formed interfaces, which can decrease the dissociation barrier of O2, improve the reduction of the adsorbed oxygen, and increase the oxidation state of the interfacial Mn ion [63].
Figure 12
Beside compositing with other catalysts, the surface acidity and alkalinity of SmMn2O5 can also be modulated by compositing with other acidic or basic materials, benefiting for the effective adsorption of gas molecules. Chen et al. [94] increased the surface alkalinity of the SmMn2O5 by complexing appropriate amounts of K-based oxide (K2Mn4O8) for NOx adsorption and further reaction. This enhanced adsorption capacity helps *NO2 to participate in the oxidation directly and accelerates the reaction thoroughly. As shown in Fig. 12d, 5% KMO-SMO sample has showed the best activity for soot combustion, with T50 of 336 ℃. Similarly, the modification of the amorphous acid oxide Nb2O5 on the SmMn2O5 catalyst increases the number of -NH groups on the surface, which is suitable for the catalytic oxidation of the basic gas NH3 and further improves its reaction selectivity to N2 [69]. In conclusion, the acid-base complementary design of the catalyst according to the acidity and alkalinity of different gasses is conducive to improve the inherent activity of the catalyst for specific reactions and release the maximum catalytic ability of SmMn2O5.
On the other hand, the semiconductor properties and poor electronic conductivity of SmMn2O5 also limit its catalytic ability. Compounding SmMn2O5 with a highly conductive substrate is one of the effective ways to improve its conductivity. SmMn2O5 nanoparticles was successful synthesized on nitrogen-doped reduced graphene oxide (SMO@NrGO) for highly efficient ORR [95]. A comparable half-wave potential of 0.84 V to that of the Pt/C was achieved by efficiently storing and compensating electrons at the interfacial Mn-N(C) bonds of SMO@NrGO (Fig. 12e). The improvement of electronic conductivity can enhance interactions between intermediates and active sites in SMO@NrGO for better catalytic performance. Beside, Chen et al. [94] found that the good thermal conductivity of carbon nanotubes (CNTs) and graphene oxide (GO) could improve the catalytic activity of SmMn2O5 for NO oxidation over a wider temperature range, attributing to the higher heat release of SmMn2O5/CNTs (200–280 ℃) and SmMn2O5/GO (250–300 ℃) in different temperature ranges (Fig. 12f). We can see that SmMn2O5 can obtain the better charge carrier transmission with the help of other highly conductive materials to achieve superior catalytic performance.
6. Conclusion and outlook
We herein presented an overview of the basic physiochemical properties, catalytic activity origin, reaction mechanism, catalytic applications and modification methods of the mullite-type oxide SmMn2O5. Due to the existence of two crystal fields, MnO6 octahedral and MnO5 pyramid, as well as the unique structure of Mn-Mn dimer, SmMn2O5 exhibits good thermal stability, ferroelectric and semiconductor properties, and good oxygen conductivity, providing the basis for its extensive catalytic applications such as automobile catalytic converters, potentiometric gas sensor, oxygen evolution reaction, supercapacitor and metal-air batteries. In order to meet the needs of various catalytic fields, the catalytic performance of SmMn2O5 oxide was further explored by regulating morphology, doping metal ions, introducing oxygen vacancies, and compounding with other materials.
Despite several years of extensive research on the catalysis and properties enhancement, much room remains to further advance the concepts reviewed herein. Although the SmMn2O5 was confirmed sensitive to many gasses, expanding the catalytic application field of SmMn2O5 is also an interesting work: (1) SmMn2O5 as a catalyst can be tried for more reactions, such as the conversion and utilization of CO2, and the catalytic reforming of methane, combining with current development needs or the latest policies. It can also be used as an electrode in various energy storage and conversion devices, such as solid oxide fuel cells and solid oxide electrolysis cell, arising from its significant catalytic activity at high temperature; (2) In order to explore the greater catalytic ability of SmMn2O5, new synthesis methods and processing conditions can be developed for the structure and morphology modulation. (3) The following strategies may be of great importance for further properties study: synchrotron radiation devices and in-situ test to study the chemical and structural changes on the surface under catalytic transients, which can provide new insights for revealing surface chemistry and catalytic mechanisms. In addition, it is more practical to introduce real reaction conditions such as temperature and pressure into theoretical simulations to study the reaction kinetics on the surface of the catalyst. We believe that a comprehensive understanding of the crystal and electronic structure, basic properties, and reaction mechanisms of mullite-type oxide SmMn2O5 will help to develop SmMn2O5 catalysts with high activity, high selectivity and long-term stability to satisfy energy and environmental needs.
Declaration of competing interest
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Acknowledgments
This research was financially supported by the National Natural Science Foundation of China (Nos. 52072134, U1910209, 51972128, 52272205), and Hubei Province (Nos. 2021BCA149, 2021CFA072, 2022BAA087).
-
-
[1]
J. Hwang, R.R. Rao, L. Giordano, et al., Science 358 (2017) 751–756. doi: 10.1126/science.aam7092
-
[2]
R.M. Bullock, J.G. Chen, L. Gagliardi, et al., Science 369 (2020) eabc3183. doi: 10.1126/science.abc3183
-
[3]
H.A. Gasteiger, N.M. Markovic, Science 324 (2009) 48–49. doi: 10.1126/science.1172083
-
[4]
R.E. Smalley, MRS Bull. 30 (2005) 412–417. doi: 10.1557/mrs2005.124
-
[5]
P.C.K. Vesborg, T.F. Jaramillo, RSC Adv. 2 (2012) 7933. doi: 10.1039/c2ra20839c
-
[6]
Z. Chen, X. Liu, K. Cho, et al., ACS Catal. 5 (2015) 4913–4926. doi: 10.1021/acscatal.5b00249
-
[7]
K.J. May, C.E. Carlton, K.A. Stoerzinger, et al., J. Phys. Chem. Lett. 3 (2012) 3264–3270. doi: 10.1021/jz301414z
-
[8]
M. Risch, A. Grimaud, K.J. May, et al., J. Phys. Chem. C 117 (2013) 8628–8635. doi: 10.1021/jp3126768
-
[9]
C.H. Kim, G. Qi, K. Dahlberg, et al., Science 327 (2010) 1624–1627. doi: 10.1126/science.1184087
-
[10]
W. Wang, G. McCool, N. Kapur, et al., Science 337 (2012) 832–835. doi: 10.1126/science.1225091
-
[11]
Y. Li, X. Zhang, H. Li, et al., Nano Energy 27 (2016) 8–16. doi: 10.1016/j.nanoen.2016.06.033
-
[12]
J.A. Alonso, M.T. Casais, M.J. Martinez-Lope, et al., J. Solid State Chem. 129 (1997) 105–112. doi: 10.1006/jssc.1996.7237
-
[13]
J.A. Alonso, M.T. Casais, M.J. Martinez-Lope, et al., J. Phys. Condens. Matter. 9 (1997) 8515–8526. doi: 10.1088/0953-8984/9/40/017
-
[14]
I. Kagomiya, K. Kohn, T. Uchiyama, Ferroelectrics 280 (2002) 131–143. doi: 10.1080/00150190214799
-
[15]
G. Zhu, P. Liu, M. Hojamberdiev, et al., Mater. Chem. Phys. 118 (2009) 467–472. doi: 10.1016/j.matchemphys.2009.08.019
-
[16]
C. Dong, Z. Liu, J. Liu, et al., Small 13 (2017) 1603903. doi: 10.1002/smll.201603903
-
[17]
R.X. Fischer, H. Schneider, The mullite-type family of crystal structures, Mullite R H. Schneider, S. Komarneni (Eds.), E-Publishing Inc., Germany, 2005, pp. 16–18.
-
[18]
T. Fujita, K. Kohn, Ferroelectrics 219 (1998) 155–160. doi: 10.1080/00150199808213511
-
[19]
M. Tachibana, K. Akiyama, H. Kawaji, et al., Phys. Rev. B 72 (2005) 224425. doi: 10.1103/PhysRevB.72.224425
-
[20]
Y. Noda, H. Kimura, M. Fukunaga, et al., J. Phys. Condens. Matter. 20 (2008) 434206. doi: 10.1088/0953-8984/20/43/434206
-
[21]
G. Yahia, F. Damay, S. Chattopadhyay, et al., Phys. Rev. B 95 (2017) 184112. doi: 10.1103/PhysRevB.95.184112
-
[22]
Y. Ishii, S. Horio, H. Yamamoto, et al., Phys. Rev. B 98 (2018) 174428. doi: 10.1103/PhysRevB.98.174428
-
[23]
S. Mansouri, S. Jandl, M. Balli, et al., Phys. Rev. B 100 (2019) 085147. doi: 10.1103/PhysRevB.100.085147
-
[24]
T. Hsu, C. Yang, C. Chu, et al., Chin. J. Phys. 62 (2019) 368–373. doi: 10.1016/j.cjph.2019.10.012
-
[25]
Y. Yun, L. Kampschulte, M. Li, et al., J. Phys. Chem. C 111 (2007) 13951–13956. doi: 10.1021/jp074214f
-
[26]
L. Wang, A. Teleki, S.E. Pratsinis, et al., Chem. Mater. 20 (2008) 4794–4796. doi: 10.1021/cm800761e
-
[27]
J. Ahmad, S.H. Bukhari, J.A. Khan, et al., Phys. Scr. 95 (2020) 115803. doi: 10.1088/1402-4896/abbcf5
-
[28]
H. Li, Z. Yang, J. Liu, et al., Appl. Phys. Lett. 109 (2016) 211903. doi: 10.1063/1.4968786
-
[29]
C. Zhao, M. Yu, Z. Yang, et al., Nano Energy 51 (2018) 91–101. doi: 10.1016/j.nanoen.2018.06.039
-
[30]
S. Chen, H. Li, Y. Hao, et al., Catal. Sci. Technol. 10 (2020) 1941–1951. doi: 10.1039/C9CY02522G
-
[31]
A.M. Yankin, O.M. Fedorova, V.F. Balakirev, et al., Russ. J. Phys. Chem. A 81 (2007) 139–142. doi: 10.1134/S0036024407010244
-
[32]
A.M. Yankin, V.F. Balakirev, O.M. Fedorova, et al., Russ. J. Phys. Chem. 80 (2006) 1714–1716. doi: 10.1134/S0036024406110033
-
[33]
C. Li, S. Thampy, Y. Zheng, et al., J. Phys. Condens. Matter 28 (2016) 125602. doi: 10.1088/0953-8984/28/12/125602
-
[34]
P. Zhao, P. Yu, Z. Feng, et al., RSC Adv. 6 (2016) 65950–65959. doi: 10.1039/C6RA12313A
-
[35]
J.K. Nørskov, F. Abild-Pedersen, F. Studt, et al., PNAS 108 (2011) 937–943. doi: 10.1073/pnas.1006652108
-
[36]
H. Li, W. Wang, X. Qian, et al., Catal. Sci. Technol. 6 (2016) 3971–3975. doi: 10.1039/C5CY01798J
-
[37]
Y. Zheng, S. Thampy, N. Ashburn, et al., J. Am. Chem. Soc. 141 (2019) 10722–10728. doi: 10.1021/jacs.9b03334
-
[38]
L. Wang, H. Li, J. Liu, et al., J. Mater. Chem. A 9 (2021) 380–389. doi: 10.1039/D0TA09537K
-
[39]
J. Liu, M. Yu, X. Wang, et al., J. Mater. Chem. A 5 (2017) 20922–20931. doi: 10.1039/C7TA02905E
-
[40]
W. Ding, X. Gu, H. Su, et al., J. Phys. Chem. C 118 (2014) 12216–12223. doi: 10.1021/jp503745c
-
[41]
B. Shan, Y. Zhao, J. Hyun, et al., J. Phys. Chem. C 113 (2009) 6088–6092. doi: 10.1021/jp8094962
-
[42]
L. Lin, P. Shi, L. Yao, et al., Nanotechnology 33 (2022) 205504. doi: 10.1088/1361-6528/ac4f19
-
[43]
D.C. Sesu, P. Marbaniang, S. Ingavale, A.C. Manohar, B. Kakade, ChemistrySelect 5 (2020) 306–311. doi: 10.1002/slct.201904127
-
[44]
P. Mars, D.W. van Krevelen, Chem. Eng. Sci. 3 (1954) 41–59. doi: 10.1016/S0009-2509(54)80005-4
-
[45]
Y. Zhang, G. Qin, J. Zheng, et al., Mol. Catal. 540 (2023) 113057. doi: 10.1016/j.mcat.2023.113057
-
[46]
Y. Hosono, H. Saito, T. Higo, et al., J. Phys. Chem. C 125 (2021) 11411–11418. doi: 10.1021/acs.jpcc.1c02855
-
[47]
C. Wang, X. Gu, H. Yan, et al., ACS Catal. 7 (2017) 887–891. doi: 10.1021/acscatal.6b02685
-
[48]
M. Zhao, M. Shen, J. Wang, J. Catal. 248 (2007) 258–267. doi: 10.1016/j.jcat.2007.03.005
-
[49]
F. Chen, D. Liu, J. Zhang, et al., J. Phys. Chem. Chem. Phys. 14 (2012) 16573–16580. doi: 10.1039/c2cp41281k
-
[50]
S. Thampy, N. Ashburn, K. Cho, et al., Adv. Energy Sustain. Res. 2 (2021) 2000075. doi: 10.1002/aesr.202000075
-
[51]
J. Rossmeisl, Z.W. Qu, H. Zhu, et al., J. Electroanal. Chem. 607 (2007) 83–89. doi: 10.1016/j.jelechem.2006.11.008
-
[52]
I.C. Man, H. Su, F. Calle-Vallejo, et al., ChemCatChem 3 (2011) 1159–1165. doi: 10.1002/cctc.201000397
-
[53]
X. Rong, J. Parolin, A.M. Kolpak, ACS Catal. 6 (2016) 1153–1158. doi: 10.1021/acscatal.5b02432
-
[54]
J.T. Mefford, X. Rong, A.M. Abakumov, et al., Nat. Commun. 7 (2016) 11053. doi: 10.1038/ncomms11053
-
[55]
A. Grimaud, W.T. Hong, Y. Shao-Horn, J.M. Tarascon, Nat. Mater. 15 (2016) 121–126. doi: 10.1038/nmat4551
-
[56]
J. Hwang, R.R. Rao, L. Giordano, et al., Science 358 (2017) 751–756. doi: 10.1126/science.aam7092
-
[57]
M.V. Twigg, Catal. Today 117 (2006) 407–418. doi: 10.1016/j.cattod.2006.06.044
-
[58]
A. Russell, W.S. Epling, Catal. Rev. Sci. Eng. 53 (2011) 337–423. doi: 10.1080/01614940.2011.596429
-
[59]
M. Piumetti, S. Bensaid, D. Fino, et al., Catal. Struct. React. 1 (2015) 155–173.
-
[60]
C. Myung, J. Kim, K. Choi, et al., Fuel 94 (2012) 348–355. doi: 10.1016/j.fuel.2011.10.041
-
[61]
R. Li, Y. Rao, Y. Huang, Chin. Chem. Lett. 34 (2023) 108000. doi: 10.1016/j.cclet.2022.108000
-
[62]
S. Thampy, Y. Zheng, S. Dillon, et al., Catal. Today 310 (2018) 195–201. doi: 10.1016/j.cattod.2017.05.008
-
[63]
B. Jin, B. Zhao, S. Liu, et al., Appl. Catal. B: Environ. 273 (2020) 119058. doi: 10.1016/j.apcatb.2020.119058
-
[64]
Z. Feng, Q. Liu, Y. Chen, et al., Catal. Sci. Technol. 7 (2017) 838–847. doi: 10.1039/C6CY02478E
-
[65]
Y. Chen, C. Du, Y. Lang, et al., Catal. Sci. Technol. 8 (2018) 5955–5962. doi: 10.1039/C8CY01663A
-
[66]
X. Feng, R. Liu, S. Zhang, et al., Chem. Afr. 3 (2020) 695–701. doi: 10.1007/s42250-020-00136-5
-
[67]
Y. Chen, X. Chen, X. Ma, et al., J. Catal. 402 (2021) 10–21. doi: 10.1016/j.jcat.2021.07.027
-
[68]
X. Wan, L. Wang, S. Gao, et al., Chem. Eng. J. 410 (2021) 128305. doi: 10.1016/j.cej.2020.128305
-
[69]
R. Liu, B. Zhou, L. Liu, et al., J. Colloid Interface Sci. 585 (2021) 302–311. doi: 10.1016/j.jcis.2020.11.096
-
[70]
Z. Feng, C. Du, Y. Chen, et al., Catal. Sci. Technol. 8 (2018) 3785–3794. doi: 10.1039/C8CY00897C
-
[71]
W. Li, H. Mao, B. Jin, et al., Fuel 306 (2021) 121685. doi: 10.1016/j.fuel.2021.121685
-
[72]
Y. Zhu, C. Du, Z. Feng, et al., RSC Adv. 8 (2018) 5459–5467. doi: 10.1039/C7RA11551B
-
[73]
Q. Yang, X. Wang, X. Wang, et al., ACS Catal. 11 (2021) 14507–14520. doi: 10.1021/acscatal.1c03955
-
[74]
B.J. Rani, G. Ravi, R. Yuvakkumar, et al., Vacuum 166 (2019) 279–285. doi: 10.1016/j.vacuum.2019.05.029
-
[75]
B.J. Rani, M. Gowsalya, G. Ravi, et al., Mater. Res. Express 6 (2019) 95090. doi: 10.1088/2053-1591/ab3333
-
[76]
Z. Zhu, L. Zheng, S. Zheng, et al., Ceram. Int. 45 (2019) 885–891. doi: 10.1016/j.ceramint.2018.09.260
-
[77]
B. Yang, J. Xu, C. Wang, et al., Mater. Chem. Phys. 245 (2020) 122679. doi: 10.1016/j.matchemphys.2020.122679
-
[78]
H. Kim, J. Lee, Sens. Actuators. B: Chem. 192 (2014) 607–627. doi: 10.1016/j.snb.2013.11.005
-
[79]
J.W. Fergus, Sens. Actuators. B: Chem. 123 (2007) 1169–1179. doi: 10.1016/j.snb.2006.10.051
-
[80]
B. Liao, Q. Wei, K. Wang, et al., Sens. Actuators. B: Chem. 80 (2001) 208–214. doi: 10.1016/S0925-4005(01)00892-9
-
[81]
F. Chu, C. Zuo, Z. Tian, et al., J. Alloy. Compd. 748 (2018) 375–381. doi: 10.1016/j.jallcom.2018.03.166
-
[82]
M. Yu, Q. Wei, M. Wu, et al., J. Power Sources 396 (2018) 754–763. doi: 10.1016/j.jpowsour.2018.06.095
-
[83]
S. Thampy, V. Ibarra, Y. Lee, et al., Appl. Surf. Sci. 385 (2016) 490–497. doi: 10.1016/j.apsusc.2016.05.151
-
[84]
S. Ma, X. Wang, T. Chen, et al., Chem. Eng. J. 354 (2018) 191–196. doi: 10.1016/j.cej.2018.07.197
-
[85]
X. Wang, T. Chen, Y. Zhang, et al., Mol. Catal. 516 (2021) 111983. doi: 10.1016/j.mcat.2021.111983
-
[86]
Z. Feng, J. Wang, X. Liu, et al., Catal. Sci. Technol. 6 (2016) 5580–5589. doi: 10.1039/C5CY01919B
-
[87]
J. Yang, J. Zhang, X. Liu, et al., J. Catal. 359 (2018) 122–129. doi: 10.1016/j.jcat.2018.01.002
-
[88]
S. Thampy, N. Ashburn, S. Dillon, et al., J. Phys. Chem. C 124 (2020) 15913–15919. doi: 10.1021/acs.jpcc.0c03443
-
[89]
X. Zhao, L. Wang, X. Chen, et al., J. Power Sources 449 (2020) 227482. doi: 10.1016/j.jpowsour.2019.227482
-
[90]
X. Liu, Y. Tang, M. Shen, et al., Chem. Sci. 9 (2018) 2469–2473. doi: 10.1039/C7SC05486F
-
[91]
Y. Lang, J. Zhang, Z. Feng, et al., Catal. Sci. Technol. 8 (2018) 5490–5497. doi: 10.1039/C8CY01263F
-
[92]
X. Liu, J. Yang, G. Shen, et al., Nanoscale 11 (2019) 8150–8159. doi: 10.1039/C8NR09054H
-
[93]
Y. Chen, G. Shen, Y. Lang, et al., J. Catal. 384 (2020) 96–105. doi: 10.1016/j.jcat.2020.02.006
-
[94]
T. Chen, X. Wang, S. Ma, et al., Solid State Sci. 108 (2020) 106425. doi: 10.1016/j.solidstatesciences.2020.106425
-
[95]
M. Yu, L. Wang, J. Liu, et al., ACS Appl. Mater. Interfaces 11 (2019) 17482–17490. doi: 10.1021/acsami.9b04451
-
[96]
L. Gao, X. Zhong, J. Chen, et al., Chin. Chem. Lett. 34 (2023) 108085. doi: 10.1016/j.cclet.2022.108085
-
[97]
H. Wang, T. Zhu, Y. Qiao, S. Dong, Z. Qu, Chin. Chem. Lett. 33 (2022) 5223–5227. doi: 10.1016/j.cclet.2022.01.075
-
[98]
B. Xia, G. Wang, S. Cui, et al., Chin. Chem. Lett. 34 (2023) 107810. doi: 10.1016/j.cclet.2022.107810
-
[99]
Z. Cirena, Y. Nie, Y. Li, et al., Chin. Chem. Lett. 34 (2023) 107726. doi: 10.1016/j.cclet.2022.08.006
-
[1]
-
Figure 2 (a) Crystal structure of mullite SmMn2O5. The dashed circles denote two O sites with the lowest oxygen vacancy formation energy, named O3 and O2, respectively. Reproduced with permission [16]. Copyright 2017, Wiley-VCH. (b) Schematic view of the conduction mechanism of bridging oxygen atoms (Obri) in SmMn2O5. Reproduced with permission [17]. Copyright 2005, Wiley-VCH.
Figure 3 (a) Variation of dielectric constant (left) and loss tangent (right) as a function of frequency at various temperatures. Reproduced with permission [27]. Copyright 2020, IOP Science. (b) Change of resistivity as a function of temperature. Reproduced with permission [27]. Copyright 2020, IOP Science. (c) Variation of conductivity as a function of frequency. Reproduced with permission [27]. Copyright 2020, IOP Science. (d) Temperature dependence of the oxygen partial pressure for the thermal dissociation of SmMn2O5. Reproduced with permission [31]. Copyright 2007, Springer. (e) Thermal stability diagram of SmMn2O5 compounds. Reproduced with permission [33]. Copyright 2016, IOP Science. (f) XRD profiles of the SmMn2O5 (SMO) samples. Reproduced with permission [34]. Copyright 2016, Royal Society of Chemistry.
Figure 4 (a) The Mn-O octahedral and pyramidal crystalline fields, and the corresponding the d-orbital splitting configurations. Reproduced with permission [29]. Copyright 2018, Elsevier. (b) The pd hybridization strength versus the NO conversion rate. Reproduced with permission [36]. Copyright 2016, Royal Society of Chemistry. (c) Schematic diagram of the connection between the Mnoct-3d and Obulk-2p hybridization and the desorption of O*. The energy of the dz2 orbital could be modulated by the hybridization between Mnoct-3d and Obulk-2p, which consequently results in tunable bonding strength between Mnpyr and O*. Reproduced with permission [36]. Copyright 2016, Royal Society of Chemistry.
Figure 5 Schematic of NO oxidation reaction pathways with mechanisms of (a) L-H, (b) ER, (c) classic MvK, (d) cooperative MvK, Reproduced with permission [50]. Copyright 2021, Wiley-VCH. and (e) coexistence of MvK and ER. The green balls, red balls, and red dashed circles represent N atoms, O atoms, and oxygen vacancies, respectively.
Figure 6 ORR mechanism diagrams with a comparison between (a) AEM, (b) LOM, and (c) LAM. The atoms marked in red represent the original atoms of the system, while the adsorbed substances are labeled in black. Reproduced with permission [38]. Copyright 2021, Royal Society of Chemistry.
Figure 7 (a) NO, (b) CO, (c) soot (d) VOC and (e) HC oxidation reactions catalyzed by SmMn2O5 oxide. Top: Schematics of various oxidation reactions in (a–e). The green, black, white, and red balls represent N, C, H, and O atoms, respectively. Middle: Comparison of light-off temperature (T50) for SmMn2O5 (red) and commercial Pt/Al2O3 (blue) for the oxidation reactions in (a–e). Data adapted from Refs. [10,34,64,30,71]. Bottom: Catalytic activities of SmMn2O5 for the oxidation reactions in (a–e). Reproduced with permission [10,34,64,30,71]. Copyright 2012, American Association for the Advancement of Science; Copyright 2016, Royal Society of Chemistry; Copyright 2017, Royal Society of Chemistry; Copyright 2020, Royal Society of Chemistry; Copyright 2021, Elsevier.
Figure 8 (a) CV, (b) EIS, (c) GCD, (d) LSV study, (e) Tafel slope and (f) CA study of mullite SmMn2O5 Hollow long cuboids. Reproduced with permission [74]. Copyright 2019, Elsevier.
Figure 9 (a) Schematic diagram of the prepared sensor. Reproduced with permission [77]. Copyright 2020, Elsevier. (b) Response and recovery time of SmMn2O5 sample to 100 ppm acetone at 300 ℃. Reproduced with permission [76]. Copyright 2019, Elsevier. (c) long-term stability test of the sensor attached with SmMn2O5 sensing electrode. Reproduced with permission [77]. Copyright 2020, Elsevier.
Figure 10 (a) Schematic of a primary Mg-air battery using mixed-mullite catalyst in oxygen cathode, Mg anode, and aqueous 1 mol/L NaCl electrolyte. Reproduced with permission [11]. Copyright 2016, Elsevier. (b) LSV curves of oxygen reduction on pure SmMn2O5, carbon, MnOx/C, SmMn2O5-NRs/C and Pt/C in O2-saturated 0.1 M KOH at a rotating speed of 1600 rpm with a scan rate of 5 mV/s. Reproduced with permission [39]. Copyright 2017, Royal Society of Chemistry. (c) Chronoamperometric curves of SmMn2O5-NRs/C and Pt/C at a constant potential of 0.6 V vs. the RHE. Reproduced with permission [39]. Copyright 2017, Royal Society of Chemistry. (d) Typical polarization curves and the corresponding power density plots of the batteries made with SmMn2O5-NRs and Pt/C. Reproduced with permission [82]. Copyright 2018, Elsevier. (e) Cycling test of the ZAB made with SmMn2O5-NRs at different charging-discharging current density. Reproduced with permission [82]. Copyright 2018, Elsevier.
Figure 11 (a) SEM images of SmMn2O5 samples. Reproduced with permission [64,71,84,85]. Copyright 2017, Royal Society of Chemistry; Copyright 2021, Elsevier; Copyright 2018, Elsevier; Copyright 2021, Elsevier. (b) The schematic diagram of the interface effect: mullite (SmMn2O5)/O-deficient perovskite (BaMnO2.83). Reproduced with permission [29]. Copyright 2018, Elsevier. (c) Specific and mass activities at the potential of 0.9 V (vs. RHE). Reproduced with permission [29]. Copyright 2018, Elsevier. (d) Energetic routes of O2 adsorption and dissociation on the Mn2 dimer and Mn-X heterodimers at the interface of Pt20/SMO. The corresponding barrier energies of O2 dissociation have been labeled. Reproduced with permission [92]. Copyright 2019, Royal Society of Chemistry. (e) Specific activities (left) and mass activities (right) as a function of Mn valence state and δ value in SMO5-δ at 0.63 V (vs. RHE). Reproduced with permission [89]. Copyright 2020, Elsevier.
Figure 12 (a) A comparison chart of the performance improvement of different materials composited with SmMn2O5 (SMO). PGMs, MOx, C and AB represent platinum group metals, metal oxides, the highly conductive material, and acidic or basic materials, respectively. (b) CO conversion of Pt/SMO, PtIWI/SMO, Pt/Al2O3 and SMO as a function of reaction temperature. Reproduced with permission [90]. Copyright 2018, Royal Society of Chemistry. (c) Apparent activation energy Ea of the MOx/SMO samples. Reproduced with permission [91]. Copyright 2018, Royal Society of Chemistry. (d) Soot combustion with NOx-assisted. Reproduced with permission [93]. Copyright 2020, Elsevier. (e) Half-wave potential and Tafel’s slope of SMO, NrGO, SMO/NrGO-2, SMO@rGO-2, SMO@NrGO-2, and Pt/C. Reproduced with permission [95]. Copyright 2019, American Chemical Society. (f) Differential scanning calorimetry (DSC) curves for SmMn2O5 nanocrystals and their composites. Reproduced with permission [94]. Copyright 2020, Elsevier.
计量
- PDF下载量: 1
- 文章访问数: 299
- HTML全文浏览量: 4