无机材料学报, 2021, 36 (1): 88, 网络出版: 2021-01-21  

铋掺杂提高氧化铈中氧空位浓度增强CO2光催化还原性能

Bi-doped Ceria with Increased Oxygen Vacancy for Enhanced CO2 Photoreduction Performance
刘亚鑫 1,2王敏 1,2沈梦 1,2王强 1,2张玲霞 1,2,*
作者单位
1 中国科学院 上海硅酸盐研究所, 高性能陶瓷和超微结构国家重点实验室, 上海 200050
2 中国科学院大学 材料科学与光电技术学院, 北京 100049
摘要
氧空位在CO2光催化还原过程中往往发挥重要作用。本工作中, 用水热法合成了不同Bi掺杂量的二氧化铈光催化剂Ce1-xBixO2-δ, 其中Ce0.6Bi0.4O2-δ在Xe灯照射下表现出最高的光催化活性, 其CO产率为纯二氧化铈纳米棒的4.6倍。X射线衍射(XRD)分析表明固溶体保留了二氧化铈的萤石结构;紫外-可见漫反射(UV-Vis)光谱表明固溶体可见光吸收增强;X射线光电子能谱 (XPS)和拉曼光谱(Raman)分析表明, 掺杂后氧空位浓度明显提高。结合原位傅里叶变换红外光谱(in-situ FT-IR), 发现引入Bi提高了固溶体中氧空位的浓度, 并改变了CO2在催化剂表面上的吸附/活化行为, 光照下碳酸氢根、碳酸根、甲酸等中间产物明显增多, 从而增强了CO2光催化还原性能。
Abstract
Oxygen vacancy plays an important role in promoting CO2 adsorption and reduction on photocatalysts. Bi was heavily doped into ceria, forming a solid solution catalyst Ce1-xBixO2-δ meanwhile maintaining the fluorite structure, to increase the oxygen vacancy concentration. The sample Ce0.6Bi0.4O2-δ showed the highest photocatalytic activity with a CO yield of ~4.6 times that of the pristine ceria nanorods. Bi was homogeneously dispersed into the fluorite ceria which was confirmed by XRD and EDX elemental mapping. It has been evidenced by the results of Raman and XPS that Bi introduction boosts the concentration of oxygen vacancy in the solid solution that can facilitate the adsorption/activation of carbonate and bicarbonate intermediates on its surface according to in-situ FT-IR.

Solar energy driving CO2 conversion into valuable chemicals is a promising way to tackle the dilemma of climate crisis and energy shortage[1]. Since the discovery that TiO2 can be used to reduce H2O and CO2, a variety of semiconductor photocatalysts have been fabricated to increase the photo-to-chemical efficiency. Based on these materials, a series of improvement strategies have been developed, such as defect engineering[2], surface plasmon sensitization[3], element doping, and heterojunction construction[4]. Most strategies gain improved photocatalytic performance by increasing visible light absorption[5], promoting charge separation/transfer and enhancing CO2 adsorption/activation.

Among the recent reports on metal oxide based photocatalysts, oxygen vacancy (Ov) engineering has been found to be a facile and effective way to improve photoactivity on all respects. Formation of Ov in Bi2MoO6 created a defect energy level in the band gap and enhanced the visible-light-driven photocatalytic activity of Bi2MoO6[6]. Our previous work introduced Cu into CeO2-x to generate and stabilize the surface oxygen vacancies (Ovs) and achieved high CO2 photoreduction activity[7]. Though researchers have made great efforts to explore the effects of Ov on catalysts’ chemical/physical properties and photocatalytic reaction process by experiments and theoretical calculations, there are still some issues to be clarified based on rational material design and advanced characterizations, especially in-situ or operando investigations.

Typically, ceria-based oxides serve as an active oxygen donor in a wide range of reactions[8]. It is worth noting that both Ce3+ and Ce4+ can exist stably because of their unique electronic configurations. To balance the charge, active oxygen is released spontaneously and Ov is formed, which can be usually recovered when exposed to an oxidative environment. Due to this unique nature, great efforts have been made to create Ov and clarify the effects of Ov in CeO2. In order to demonstrate the effect and behavior of the as-formed Ov in CeO2, bismuth is selected as a dopant that is expected to maintain the fluorite structure and the stability of Ov because of its lower valence state (+3) and its ion radius (0.117 nm) similar to Ce4+ (0.097 nm)[9]. Here we synthesized and explored a Ce1-xBixO2-δ solid solution photocatalyst with boosted Ov concentration and enhanced CO2 photoreduction ability under ambient environment. A series of Ce1-xBixO2-δ oxide catalysts were investigated and the catalyst with an optimal Bi/(Bi+Ce) ratio of 0.4 exhibited highest CO2 photoreduction efficiency. The relationship between the Ov concentration and the photocatalytic activity of Ce1-xBixO2-δ solid solution catalyst was explored by structural, compositional and in-situ FT-IR characterizations.

1 Experimental

1.1 Preparation of catalysts

Samples were synthesized by a facile one-step hydrothermal process at controlled conditions. Typically, a certain amount of Ce(NO3)3·6H2O and Bi(NO3)3·5H2O were dissolved in diluted nitrite acid (HNO3, 5 mL, 4 mol/L) by sonication before the dropwise addition of 8 mol/L aqueous solution of NaOH (35 mL) while stirring. This mixture was kept stirring for 30 min till the formation of a gelatinous suspension and then transferred into a 50 mL Teflon-lined autoclave within a stainless-steel tank and placed at 100 ℃ for 24 h. The obtained samples were rinsed in deionized water and washed with anhydrous ethanol for several times. Then the powder was oven-dried overnight in vacuum at 60 ℃. The samples were denoted as Ce1-xBixO2-δ, where x represents the molar ratio of Bi/(Ce+Bi). The sample synthesized with Ce(NO3)3·5H2O (1.5 g) as the sole precursor under the same conditions was named as CeO2.

1.2 Characterizations

X-ray diffraction patterns (XRD) of the samples were recorded on Rigaku D/Max 2200PC X-ray diffractometer. The measurement was operated at room temperature under Cu Kα radiation with a scanning rate of 4 (°)/min. A JEM-2100F field emission transmission electron microscope (TEM) (200 kV) was used to obtain the TEM image. Ultraviolet visible (UV-Vis) absorption spectra (800 nm to 200 nm) were recorded by a UV-3101 PC Shimadzu spectroscope (BaSO4 as the reference standard material). Electron paramagnetic resonance (EPR) spectra were obtained from a Bruker EMXplus (Germany) spectrometer at 90 K. X-ray photoelectron spectra (XPS) were obtained on a Thermo Scientific ESCALAB 250 spectrometer with multichannel detector, Al Kα radiation as excitation source, and C1s at 284.6 eV as a signal- calibrating standard of binding-energy values. Time-resolved fluorescence spectra were tested on an Edinburgh Instruments FLS920 spectrometer under 420 nm excitation of deuterium lamp. DXR Raman microscope was used to obtain Raman spectra under the excitation of 532 nm. In-situ Fourier transform infrared (FT-IR) spectra were collected in N2 and CO2, respectively, on Nicolet iS10 equipped with MCT detector. The catalyst was swept by N2 for 1 h and stabilized for 1 h before the FT-IR spectrum was collected and saved as background data. Subsequently, the catalyst was swept by CO2 for 1 h and stabilized for 1 h. Then the in-situ FT-IR spectra were recorded at different time intervals since the beginning of light irradiation.

1.3 Electrochemical measurements

Electrochemical measurements were conducted on a CHI660A electrochemical workstation (Shanghai Chenhua, China) with a standard three-electrode system. Fluorine-doped tin oxide (FTO) glass deposited (15 mm× 25 mm) with photocatalyst as the working electrode, platinum wire and Ag/AgCl as counter electrode and reference electrode, respectively. The electrolyte was 0.2 mol/L Na2SO4 aqueous solution. The preparation process of working electrodes by electrophoretic deposition was as follows: sample powder (20 mg) and iodine (10 mg) were mixed and well milled in an agate mortar then well-dispersed in acetone (30 mL). Thus, the plating solution was obtained. A thin film of the sample was uniformly deposited on FTO with a potentiostat bias of 10 V for 10 min and then calcined in an oven at 150 ℃ for 2 h. During the electrochemical measurement, the coated area of all samples was controlled at 1 cm2. Mott-Schottky plots were obtained by impedance-potential tests at the frequency of 1000 Hz with a voltage amplitude of 10 mV. Nyquist plots were tested in dark at a bias voltage of -0.4 V vs Ag/AgCl electrode. With irradiation from a 300 W Xe arc lamp, the transient photocurrent densities were obtained at 0.3 V versus Ag/AgCl electrode.

1.4 Photocatalytic activity evaluation

The light source was a 300 W Xenon arc lamp from Aulight CEL-HX, Beijing, the light intensity of which tested on the position of catalyst was 210 mW/cm2. The gas products (CO, CH4) were analyzed by GC-2014 gas chromatograph (flame ionization detector, FID) equipped with a 0.5 nm molecular column and a TDX-01 packed column. The photocatalysis test was conducted typically as follows: 50 mg of photocatalyst was uniformly dispersed on a 2.5 cm × 2.5 cm glass at the bottom of a 625 mL sealed glass reactor. The temperature of the reactor was kept at 15 ℃ with cooling water circulation. To exclude possible influence of contaminants, the sealed reactor was filled with N2 and subjected to the irradiation of Xe lamp for 2 h before CO2 (99.99% of purity) moisturized by a water bubbler was pumped into the reactor and stabilized for 30 min. During the photoreaction, 1 mL of sample gas was continually extracted from the reactor every 1 h and analyzed on GC according to external standard method. Before each activity (or cycle) test, the photocatalyst was firstly heat-treated at 150 ℃ for 2 h, so as to remove the organic impurities and adsorbed carbon species.

2 Results and discussion

2.1 Improved photocatalytic activity of Ce1-xBixO2-δ in CO2 reduction reaction

A series of Ce1-xBixO2-δ (x=0, 1, 2, 3, 4, 5) samples were synthesized by co-precipitation method and all catalysts were tested for gas-solid phase CO2 reduction in sealed glass reactor under the conditions of 15 ℃, 1×105 Pa. CO was the main product with trace CH4 detected. Controlled experiments without light irradiation, CO2, or photocatalysts were conducted, in which no CO was detected, proving that CO was derived from CO2 photoreduction. As displayed in Fig. 1, with the increase of Bi content, the catalytic performance of the samples improves. Ce0.6Bi0.4O2-δ shows the highest activity in CO2 photoreduction with a CO yield of 0.5 μmol·g-1·h-1 under Xe light irradiation, which is ~4.6 times of that on CeO2.

10.15541/jim20200142.F001Time-dependent CO evolution over CeO<sub>2</sub>, Ce<sub>1-</sub><i><sub>x</sub></i>Bi<i><sub>x</sub></i>O<sub>2-</sub><i><sub>δ</sub></i> and Bi<sub>2</sub>O<sub>3</sub>

2.2 Chemical and structural characterizations

XRD patterns (Fig. S1) show that the Ce1-xBixO2-δ solid solution samples prepared in this work maintained the fluorite cubic structure of ceria (JCPDS 34-0394). With Bi content ranging from 10% to 50%, all peaks broaden and shift slightly towards small angle, which indicates that adding Bi into ceria increased the lattice parameter. The morphology of the as-prepared ceria and Ce0.6Bi0.4O2-δ were further characterized by TEM (Fig. 2). Pure ceria shows the morphology of granite nanorods, 100-300 nm in length and 12-20 nm in diameter (Fig. 2(a)) while Ce0.6Bi0.4O2-δ is in the form of nanocubes with particle size of ~35 nm (Fig. 2(b)). The substitution of Ce4+ atom with Bi3+ distorted the crystal structure and changed the crystal growth behavior, leading to surface reconstruction and defects, which can be confirmed by the bright spots observed in Ce0.6Bi0.4O2-δ nanocubes[10]. Fig. 2(c) shows that O, Ce, and Bi distribute homogeneously in the Ce0.6Bi0.4O2-δ sample. No segregation has been observed and a solid solution catalyst can be identified.

10.15541/jim20200142.F002TEM images of (a) CeO<sub>2</sub> and (b) Ce<sub>0.6</sub>Bi<sub>0.4</sub>O<sub>2-</sub><i><sub>δ</sub></i>, (c) EDX elemental mapping images of Ce<sub>0.6</sub>Bi<sub>0.4</sub>O<sub>2-</sub><i><sub>δ</sub></i>

To further probe the chemical and compositional differences Bi doping induced, XPS spectra were collected to analyze the chemical states of Ce, Bi and O on the surface of CeO2 and Ce0.6Bi0.4O2-δ samples. There are multiple peaks in the Ce3d XPS spectra (Fig. 3(a)). Both CeO2 and Ce0.6Bi0.4O2-δ show eight major characteristic peaks located at 917.9, 916.6, 907.2, 900.7, 898.5, 889.7, 886.5 and 882.3 eV (spin-orbit splitting peaks), which are uniquely attributed to Ce4+, thus the main valence state of Ce can be identified as +4. The presence of Ce3+ can be also distinguished in both CeO2 and Ce0.6Bi0.4O2-δ by the peaks at 902.5 and 884.2 eV(denoted as v1 and v2). Through the comparison of these two characteristic peaks (v1 and v2) corresponding to Ce3+, especially against the concurrent decrease of the two adjacent peaks(u1 and u2) of Ce4+, we are convinced of that, with the introduction of Bi into ceria lattice, Ce3+ increases in Ce0.6Bi0.4O2-δ[3]. While in Fig. 3(b), the representative core level Bi4f spectrum shows that Bi3+ exists exclusively in nanocrystalline solid solution Ce0.6Bi0.4O2-δ[10]. Two highly symmetric peaks are located at 158.8 eV for Bi 4f7/2 and 164.1 eV for Bi4f5/2, which agree well with those observed in standard bismuth oxides[11]. This result combined with those of XRD and TEM demonstrates that Ce atoms have been successfully substituted by Bi atoms in CeO2 lattice and a solid solution catalyst has been formed.

10.15541/jim20200142.F003High resolution XPS spectra of (a) Ce3d, (b) Bi4f, (c) O1s on the surface of Ce<sub>0.6</sub>Bi<sub>0.4</sub>O<sub>2-</sub><i><sub>δ</sub></i> and (d) Raman spectra of CeO<sub>2</sub> and Ce<sub>0.6</sub>Bi<sub>0.4</sub>O<sub>2-</sub><i><sub>δ</sub></i>

The O1s spectra of the samples (Fig. 3(c)) can be deconvoluted into three peaks. For CeO2, the two peaks at 529.2 and 529.7 eV are assigned to lattice oxygen species (Olatt), O2- and O22-, respectively[2]. The peak at 531.6 eV is ascribed to adsorbed oxygen species (Oad)[12,13]. It is noted that there is slight band shift of Oad to lower binding energy in the O1s spectra of Ce0.6Bi0.4O2-δ than in CeO2 and the signal increases, suggesting the adsorbed H2O and CO2 are more easily to get electrons from the surface of Ce0.6Bi0.4O2-δ, that is, are more readily to be activated on the reactive sites after the introduction of Ovs together with Bi. The Oad content calculated from peak area, which is proportional to that of Ov, is 44%, much higher in Ce0.6Bi0.4O2-δ than 34% in CeO2. This means our intention of modulating Ov content by Bi-doping is feasible.

In the Raman spectra (Fig. 3(d)), CeO2 and Ce0.6Bi0.4O2-δ exhibit distinct vibrational bands at 350-650 cm-1. For pristine CeO2, the band at 455 cm-1 is ascribed to the F2g vibration mode of CeO2 fluorite while the weak signal at about 600 cm-1 is related to the intrinsic Ovs due to the existence of Ce3+[14]. Compared with pure CeO2, major differences appear in the Raman vibration modes of Ce0.6Bi0.4O2-δ in that the F2g band shifts to low frequency (about 450 cm-1) and slightly broadens. This phenomenon is explained as that the symmetrical stretching vibration of Ce[O]8 unit is distorted by Bi doping. The second vibration mode of Ce0.6Bi0.4O2-δ located at 525 cm-1 can be indexed to the presence of Bi3+. More importantly, the signal around 585 cm-1, which is related to extrinsic Ovs that serve as charge-compensating defects, is significantly strengthened[15,16,17]. Here, it can be concluded that Bi doping in CeO2 has boosted the presence of Ovs. For Ce0.6Bi0.4O2-δ, the half-width at half-maximum (HWHM) at 450 cm-1 is ~29.8 cm-1, based on which the intrinsic Ovs concentration is calculated to be ~2.34×1021 cm-3. It is higher than that of the pristine CeO2 (2.10×1021 cm-3). This can be also verified by the enhancement of EPR signal at g-value of 2.003 corresponding to Ovs of Ce1-xBixO2-δ compared with that of CeO2 in Fig. S2. In following discussions, corresponding implications on the catalytic activity of the solid solution catalyst will be illustrated from aspects including visible light absorption, electrochemical performance and CO2 adsorption/activation.

UV-Vis diffuse absorbance spectra of all the samples were recorded as displayed in Fig. S3. Compared with pure ceria, all Ce1-xBixO2-δ solid solution samples show enhanced and red-shifted absorption. Among them, samples Ce0.6Bi0.4O2-δ and Ce0.5Bi0.5O2-δ show the strongest UV-visible light absorption that extends from 380 nm to around 500 nm. The result and the yellow color of the solid solution catalyst (pure ceria is white) indicate that the introduction of Bi can successfully extend the light response to UV-visible region, which can be attributed to the defect energy level generated from the increased Ov concentration. Based on the UV-Vis diffuse absorbance spectra and Kubelka-Munk equation[18] for indirect gap semiconductors, the bandgaps of pure ceria and Ce0.6Bi0.4O2-δ were estimated to be 3.1 and 2.6 eV, respectively. This bandgap narrowing caused by Bi incorporation and Ov can initiate the effective visible light induced photocatalytic CO2 reduction reaction.

To better understand the influence of Bi doping on the energy band structure of the solid solution catalyst, the Mott-Schottky plots of CeO2 and Ce0.6Bi0.4O2-δ electrodes were analyzed, both of which are S-type curves of n-type semiconductors (Fig. S4(a)). The flat band potentials (Vfb) were estimated from the linear part of the plots as -0.59 and -0.48 V vs Ag/AgCl electrode for CeO2 and Ce0.6Bi0.4O2-δ, respectively. According to literature, the conduction band potential (Ecb) of one n-type semiconductor is 0-0.2 V more negative than its Vfb, depending on carrier concentration and effective mass of electron. Here, the potential difference is set as 0.1 V. Based on this result, the band structure can be reasonably estimated in which a downshift of Ecb (from -0.7 V of CeO2 to -0.6 V of Ce0.6Bi0.4O2-δ) and an upshift of Evb (from 2.4 V of CeO2 to 2 V of Ce0.6Bi0.4O2-δ) are clearly illustrated. This band narrowing happens when the Bi6p and Bi6s orbits participate in the construction of CB and VB, respectively[12].

EIS Nyquist plots of the as-obtained samples in dark demonstrate that Ce0.6Bi0.4O2-δ is much more conductive than CeO2 (Fig. S4(b)). This is because that ceria is a kind of good ion conductive solid electrolyte and Ov can promote ion transportation[13] and thus endows the Ce0.6Bi0.4O2-δ solid solution catalyst better conductivity which could efficiently promote the separation and interfacial transmission of photoexcited carriers. As a result, the photocurrent response of Ce0.6Bi0.4O2-δ nearly triples over that of CeO2 (Fig. S5).

2.3 In-situ probe into the catalytic mechanism of CO2 photoreduction on Ce0.6Bi0.4O2-δ

Based on the electronic structure estimation, a possible CO2 photoreduction mechanism is proposed in Fig. 5. To further understand the origin of enhanced CO2 photocatalytic reduction activity of the solid solution catalyst, in-situ FT-IR spectroscopy was used to investigate the adsorption, activation, and conversion process of CO2 on Ce0.6Bi0.4O2-δ and CeO2. The in-situ FT-IR spectra of CO2 adsorption on CeO2 and Ce0.6Bi0.4O2-δ swept by humid CO2 and stabilized for 1 h were collected and shown in Fig. 4(a, b), respectively. Similar modes of carbonate signals after humid CO2 sweeping can be observed on the surface of the catalysts, with those of Ce0.6Bi0.4O2-δ more intense[14]. Specifically, the band at ~1700 cm-1 corresponds to bicarbonate (HCO3-)[15]. The bands at about 1510 and 1380 cm-1 are assigned to mono-dentate carbonate (m-CO32-)[16]. The bands at about 1337 and 1422 cm-1 correspond to bidentate carbonate (b-CO32-)[17]. The peaks at about 1255, 1610 and 1650 cm-1 are assigned to the bending vibration of carboxylate (COO-)[18] while the bands at about 1550 and 1200 cm-1 are ascribed to *COOH. The existence of -CO32- is due to the formation of carbonic acid and the adsorbed CO2 on the catalyst surface[2].

10.15541/jim20200142.F004<i>In-situ</i> FT-IR spectra of CO<sub>2</sub> adsorption on the surface of (a) CeO<sub>2</sub> and (b) Ce<sub>0.6</sub>Bi<sub>0.4</sub>O<sub>2-</sub><i><sub>δ</sub></i> (the spectra of the samples swept by N<sub>2</sub> for 1 h as the background); <i>In-situ</i> FT-IR spectra of CO<sub>2</sub> photoreduction on (c) CeO<sub>2</sub> and (d) Ce<sub>0.6</sub>Bi<sub>0.4</sub>O<sub>2-</sub><i><sub>δ</sub></i> at different irradiation time (The spectra collected at 0 min of irradiation in humid CO<sub>2</sub> as background)

After Xe light irradiation, the in-situ FT-IR spectra of CeO2 and Ce0.6Bi0.4O2-δ were collected and shown in Fig. 4(c, d), respectively. Over irradiation time, the clear increase and decrease of signals that is corresponding to the generation and consumption of the adsorbed intermediates on the surface of Ce0.6Bi0.4O2-δ emerges while those on CeO2 remain indiscernible. The aforementioned IR bands of both adsorbates and intermediates emerging between 1650 and 1200 cm-1 show obvious increase with prolonged irradiation time. These observations can be ascribed to the fact that the Ovs could stabilize the carbonate radicals and increase their concentration on the catalyst surface by promoting charge separation and transfer to the adsorbed intermediates[19,20]. According to literature[21], the rate limiting step of CO2 conversion to CO is the conversion of *COOH to *CO, which means the stabilized intermediates, including *COOH, cannot be readily converted to final products and explains the increase of enhanced signal of intermediate carbonates. In contrast, the peak at ~1700 cm-1 corresponding to HCO3- remains unchanged under irradiation. This indicates that bicarbonate plays a different role from other adsorbates that have signals between 1650 and 1200 cm-1. Recent reports[22,23] using operando ATR-SEIRAS discussed the effects of bicarbonate in liquid-solid interface reduction of CO2, that is, the majority of CO2 reactant near the catalyst surface originates from an equilibrium reaction with HCO3- rather than direct adsorption from the solution. Although there is still controversy in the process of CO2 reduction[24], here based on our in-situ FT-IR results, it is reasonable to propose that in a gas phase reduction of CO2 in the presence of H2O vapor, CO2 and surface adsorbed H2O form HCO3- and release active species (denoted as init- in Eq. 1) participating in CO2 reduction reaction[25]. The surface concentration of HCO3- in the equilibrium does not increase because CO2 pressure can be seen as constant. This is consistent with the photocatalytic activity test in which Ce0.6Bi0.4O2-δ shows much higher photocatalytic performance than CeO2 (Fig. 1).

CO2 (gas)+H2Oad « HCO3-ad+H+« init- (Eq. 1)

With the characteristic in-situ FT-IR bands and time related changes under irradiation, we proposed a pre-adsorption process that may shed new light into the mechanism of CO2 photoreduction (Fig. 5)[26]. It is also evidenced that the introduction of Bi has created more Ovs and formed localized electrons, which show significant benefit on the adsorption/activation behavior of CO2 on the surface of the solid solution catalyst and promote the CO2 photocatalytic activity of the catalyst.

10.15541/jim20200142.F005Scheme of the possible mechanism of CO<sub>2</sub> reduction on Ce<sub>0.6</sub>Bi<sub>0.4</sub>O<sub>2-</sub><i><sub>δ</sub></i>

3 Conclusions

In this work, Bi3+ with a lower valence and similar ion radius compared to Ce4+ has been doped into ceria forming a solid solution photocatalyst, from which it is concluded that:

1) Bi3+ doping successfully introduced Ovs into ceria while maintained its fluorite structure, thus the simultaneous introduction of defects as recombination sites of charge carriers could be effectively avoided;

2) The solid solution catalyst showed obviously improved CO2 photocatalytic reduction activity with a CO yield ~4.6 times of that on pure ceria;

3) The favorable effects of Ovs and surface localized electrons on the adsorption/activation behavior of CO2 on the catalyst surface and the separation/transfer of photoexcited carriers have been disclosed;

4) Doping aliovalent heteroatoms into metal oxide semiconductors to form solid solution catalysts is an effective way to introduce Ovs and promote photocatalytic activity.

7 Supporting materials

Supporting materials:

Bi-doped Ceria with Increased Oxygen Vacancy for Enhanced CO2 Photoreduction Performance

LIU Yaxin1,2, WANG Min1,2, SHEN Meng1,2, WANG Qiang1,2, ZHANG Lingxia1,2

1. State Key Laboratory of High Performance Ceramics and Superfine Microstructure, Shanghai Institute of Ceramics, Chinese Academy of Sciences, Shanghai 200050, China; 2. Center of Materials Science and Optoelectronics Engineering, University of Chinese Academy of Sciences, Beijing 100049, China

10.15541/jim20200142.F006XRD patterns of CeO<sub>2</sub> and Ce<sub>1-</sub><i><sub>x</sub></i>Bi<i><sub>x</sub></i>O<sub>2-</sub><i><sub>δ</sub></i>10.15541/jim20200142.F007EPR spectra of CeO<sub>2</sub> and Ce<sub>1-</sub><i><sub>x</sub></i>Bi<i><sub>x</sub></i>O<sub>2-</sub><i><sub>δ</sub></i>10.15541/jim20200142.F008UV-Vis absorption spectra of CeO<sub>2</sub> and Ce<sub>1-</sub><i><sub>x</sub></i>Bi<i><sub>x</sub></i>O<sub>2-</sub><i><sub>δ</sub></i> with inset showing the plots of (<i>αhυ</i>)<sup>2</sup><i>versus</i> photon energy10.15541/jim20200142.F009(a) Mott-Schottky plots and (b) Electrochemical impedance spectroscopy (EIS) Nyquist plots (-0.4 V (<i>vs</i> Ag/AgCl), in dark) of CeO<sub>2</sub> and Ce<sub>0.6</sub>Bi<sub>0.4</sub>O<sub>2-</sub><i><sub>δ</sub></i>10.15541/jim20200142.F010Transient photocurrent responses of CeO<sub>2</sub> and Ce<sub>0.6</sub>Bi<sub>0.4</sub>O<sub>2-</sub><i><sub>δ</sub></i>
参考文献

[1] CHANGX, WANGT, GONGJ. CO2 photo-reduction: insights into CO2 activation and reaction on surfaces of photocatalysts. Energy & Environmental Science, 2016,9(7):2177-2196.

[2] INDRAKANTI VP, KUBICKI JD, SCHOBERT HH. Photoinduced activation of CO2 on Ti-based heterogeneous catalysts: current state, chemical physics-based insights and outlook. Energy & Environmental Science, 2009,2(7):745-750.

[3] QIY, SONGL, OUYANGS, et al. Photoinduced defect engineering: enhanced photothermal catalytic performance of 2D black In2O3-x nanosheets with bifunctional oxygen vacancies. Advanced Materials, 2019,32:1903915.

[4] LINICS, CHRISTOPHERP, INGRAM DB. Plasmonic-metal nanostructures for efficient conversion of solar to chemical energy. Nature Materials, 2011,10(12):911-921.

[5] LIJ, SONG CF, PANG XJ. Controllable synthesis and photocatalytic performance of BiVO4 under visible-light irradiation. Journal of Inorganic Materials, 2019,34(2):164-172.

[6] SUNY, WANGH, XINGQ, et al. The pivotal effects of oxygen vacancy on Bi2MoO6: promoted visible light photocatalytic activity and reaction mechanism. Chinese Journal of Catalysis, 2019,40(5):647-655.

[7] WANGM, SHENM, JINX, et al. Oxygen vacancy generation and stabilization in CeO2-x by Cu introduction with improved CO2 photocatalytic reduction activity. ACS Catalysis, 2019,9(5):4573-4581.

[8] LIUY, YUS, ZHENG KW, et al. NO Photo-oxidation and in-situ DRIFTS studies on N-doped Bi2O2CO3/CdSe quantum dot composite. Journal of Inorganic Materials, 2019,34(4):425-432.

[9] JIANGD, WANGW, GAOE, et al. Bismuth-induced integration of solar energy conversion with synergistic low-temperature catalysis in Ce1-xBixO2-δ nanorods. Journal of Physical Chemistry C, 2013,117(46):24242-24249.

[10] SHAMAILAS, SAJJADA K L, CHENF, et al. Study on highly visible light active Bi2O3 loaded ordered mesoporous titania. Applied Catalysis B Environmental, 2010,94(3/4):272-280.

[11] YANG GH, MIAO WK, YUAN ZM, et al. Bi quantum dots obtained via in situ photodeposition method as a new photocatalytic CO2 reduction cocatalyst instead of noble metals: borrowing redox conversion between Bi2O3 and Bi. Applied Catalysis B Environmental, 2018,237:302-308.

[12] GAOY, LIR, CHENS, et al. Morphology-dependent interplay of reduction behaviors, oxygen vacancies and hydroxyl reactivity of CeO2 nanocrystals. Physical Chemistry Chemical Physics, 2015,17(47):31862-31871.

[13] CHEND, HED, LUJ, et al. Investigation of the role of surface lattice oxygen and bulk lattice oxygen migration of cerium-based oxygen carriers: XPS and designed H2-TPR characterization. Applied Catalysis B Environmental, 2017,218:249-259.

[14] WEBER WH, HASS KC, MCBRIDE JR. Raman study of CeO2: second-order scattering, lattice dynamics, and particle-size effects. Physical Review B: Condensed Matter, 1993,48(1):178-185.

[15] LI YF, SOHEILNIAN, GREINERM, et al. Pd@HyWO3-x nanowires efficiently catalyze the CO2 heterogeneous reduction reaction with a pronounced light effect. ACS Applied Materials & Interfaces, 2019,11(6):5610-5615.

[16] ZHUS, LIT, CAI WB, et al. CO2 Electrochemical reduction as probed through infrared spectroscopy. ACS Energy Letters, 2019,4(3):682-689.

[17] GAMARRAD, FERNANDEZ-GARCIAM, BELVERC, et al. Operando DRIFTS and XANES study of deactivating effect of CO2 on a Ce0.8Cu0.2O2. Journal of Physical Chemistry C, 2010,114(43):18576-18582.

[18] WANGY, ZHAOJ, WANGT, et al. CO2 photoreduction with H2O vapor on highly dispersed CeO2/TiO2 catalysts: surface species and their reactivity. Journal of Catalysis, 2016,337:293-302.

[19] LIJ, ZHANGW, RANM, et al. Synergistic integration of Bi metal and phosphate defects on hexagonal and monoclinic BiPO4: enhanced photocatalysis and reaction mechanism. Applied Catalysis B Environmental, 2019,243:313-321.

[20] LIX, ZHANGW, LIJ, et al. Transformation pathway and toxic intermediates inhibition of photocatalytic NO removal on designed Bi metal@defective Bi2O2SiO3. Applied Catalysis B: Environmental, 2019,241:187-195.

[21] RINGES, MORALES-GUIOC G, CHENL D,et al. Double layer charging driven carbon dioxide adsorption limits the rate of electrochemical carbon dioxide reduction on gold. Nature Communications, 2020,11(1):33.

[22] DUNWELLM, LUQ, HEYES JM, et al. The central role of bicarbonate in the electrochemical reduction of carbon dioxide on gold. Journal of the American Chemical Society, 2017,139(10):3774-3783.

[23] ZHUS, JIANGB, CAI WB, et al. Direct observation on reaction intermediates and the role of bicarbonate anions in CO2 Electrochemical reduction reaction on Cu surfaces. Journal of the American Chemical Society, 2017,139(44):15664-15667.

[24] WUTTIGA, YOONY, RYUJ, et al. Bicarbonate is not a general acid in Au-catalyzed CO2 electroreduction. Journal of the American Chemical Society, 2017,139(47):17109-17113.

[25] YEL, DENGY, WANGL, et al. Bismuth-based photocatalysts for solar photocatalytic carbon dioxide conversion. ChemSusChem, 2019,12(16):3671-3701.

[26] GRACIANIJ, MUDIYANSELAGEK, XUF, et al. Highly active copper-ceria and copper-ceria-titania catalysts for methanol synthesis from CO2. Science, 2014,345(6196):546-550.

刘亚鑫, 王敏, 沈梦, 王强, 张玲霞. 铋掺杂提高氧化铈中氧空位浓度增强CO2光催化还原性能[J]. 无机材料学报, 2021, 36(1): 88. Yaxin LIU, Min WANG, Meng SHEN, Qiang WANG, Lingxia ZHANG. Bi-doped Ceria with Increased Oxygen Vacancy for Enhanced CO2 Photoreduction Performance[J]. Journal of Inorganic Materials, 2021, 36(1): 88.

关于本站 Cookie 的使用提示

中国光学期刊网使用基于 cookie 的技术来更好地为您提供各项服务,点击此处了解我们的隐私策略。 如您需继续使用本网站,请您授权我们使用本地 cookie 来保存部分信息。
全站搜索
您最值得信赖的光电行业旗舰网络服务平台!