Читать книгу Solar-to-Chemical Conversion - Группа авторов - Страница 4

List of Illustrations

Оглавление

1 Chapter 2Figure 2.1 Schematic of solar fuel feedstocks (CO2, H2O, and solar energy) and production path on‐site and/or transported to the solar refinery [4].Figure 2.2 Schematic diagrams of (a) natural photosynthesis and (b) artificial photosynthesis based on molecular systems. Source: Liu et al. [9].Figure 2.3 (a) Traditional catalytic processeson the surface of catalysts. (b) Classic photocatalytic paths over photocatalysts.Figure 2.4 A band‐gap diagram showing the different sizes of band gaps for conductors, semiconductors, and insulators.Figure 2.5 Proposed mechanism of photocatalytic reactions on semiconductor photocatalysts. Source: Ma et al. [16].Figure 2.6 A schematic illustration of the energy correlation between semiconductor catalysts and redox couples in water. CB and VB denote a conduction band and a valence band, respectively. Source: Tu et al. [20].Figure 2.7 (a) Yields of the products produced during the photocatalytic reduction of CO2 with H2O and photoluminescence of various Ti/FSM‐16 photocatalysts. (b) Product distribution of CO2 photoreduction over 1‐TiO2, 2–10 wt% imp‐TiO2/Y‐zeolite, 3–1.0 wt% imp‐TiO2/Y‐zeolite, 4‐ex‐TiO2/Y‐zeolite, and 5‐Pt‐loaded ex‐TiO2/Y‐zeolite. (c) Schematic mechanisms of the photocatalytic CO2 reduction with H2O on TiO2. Source: (a) Ikeuea et al. [23]; (b) Anpo et al. [22]; (c) From Anpo et al. [21]. © 1995 Elsevier.Figure 2.8 Schematic cycle for the photosensitized reduction of CO2 to CH4. Source: Maidan and Willner [25].Figure 2.9 (a) Calculated band structures and total density of states of BMO‐OVs. Absorptive formats of CO2 on BMO‐OVs (b) and BMO (c). CO2 reduction pathways on BMO‐OVs (d) and BMO (e). (f) The diagram of band positions. (g) CH4 production and the TCEN value for CO2 photoreduction over the BMO‐OVs and BMO during four hours visible‐light irradiation. (h) Typical time course of CO/CH4 generated over BMO‐OVs. Source: Yang et al. [30].Figure 2.10 (a) Schematic illustration of the solar‐assisted production of methanol from CO2 and water through a three‐enzyme cascade. (b) Methanol production as a function of time in multienzymatic systems. (c) Photoelectrochemical production of methanol by TPIEC under various applied potentials under visible‐light illumination. (d) Methanol production of the TPIEC system in various multienzymatic systems with an external voltage of 0.8 V. Source: Kuk et al. [41].Figure 2.11 (a) Mechanism for the reduction of CO2 by photocatalysis under visible light with a Ru complex and an N‐Ta2O5 hybrid catalyst. (b) Turnover number for HCOOH formation from CO2 over various molecular systems as a function of irradiation time. (c) Schematic mechanism of visible‐light‐driven photocatalytic CO2 reduction over RuP/NS‐C3N4 hybrids. (d) Time courses of photocatalytic CO2 reduction using RuP/NS‐C3N4‐PU0 and RuP/NS‐C3N4‐PU90 under visible light (λ > 400 nm). Source: (a) Suzuki et al. [55]; (b) Sato et al. [55]; (c) Tsounis et al. [56]; (d) Tsounis et al. [56].Figure 2.12 (a) Scheme of the photocatalytic CO2 reduction over CdS/(Cu–NaxH2−xTi3O7) irradiated by light. (b) Gas evolution rates of C1–C3 hydrocarbons on CdS/(Cu–NaxH2−xTi3O7) as well as the specific surface area, where the Na/Ti ratios were 0.093 (low), 0.143 (medium), and 0.507 (high). (c–e) Mass spectra of the formed hydrocarbons (methane, ethane, and propane) with the labeling of 13C. (f) Proposed elementary reaction mechanisms of photocatalytic CO2 conversion into hydrocarbons. Source: Park et al. [64].Figure 2.13 Structures of CN before (a) and after (b) doping with carbon chains. (c) Electrochemical impedance spectroscopy (EIS) Nyquist plots under irradiation condition. Inset: Periodic on/off photocurrent response under visible‐light irradiation. (d) Comparison of the photocatalytic CO2 reduction rate of the samples with different amount of glycine, respectively. Source: Ren et al. [70].Figure 2.14 (a) Gas production under six hours visible‐light irradiation of pure Cu2O and Cl‐doped Cu2O samples. (b) Mass spectrum from the 18O‐labeling H2O isotope experiments of Cu2O–Cl‐4. (c) Time‐dependent product evolution over Cu2O–Cl‐4 under visible‐light irradiation. (d) Stability test of Cu2O–Cl‐4. (e) Proposed reaction pathways of CO2RR to CO and CH4 on Cl‐doped Cu2O. Source: Yu et al. [83].

2 Chapter 3Figure 3.1 The light‐dependent reactions of photosynthesis produce protons and electrons used in the conversion of CO2 to carbohydrates (CH2O) and other organic compounds, while oxidation of these organic compounds (biomass, food, fuels) by respiration or combustion releases the stored solar energy to power the metabolism of non‐photosynthetic organisms and drive human technologies.Figure 3.2 Schematic membrane arrangement of enzymes involved in the light‐dependent reactions of oxygenic photosynthesis. Red arrows indicate the flow of electrons from the primary electron donor H2O to the final electron acceptor NADP+ aided by photoexcitation in photosystems II and I. Electron flow is tied to proton translocation across the membrane. The proton gradient created by accumulation of protons in the interior of the membrane powers the synthesis of ATP.Figure 3.3 Selected types of chlorophyll molecules encountered in natural photosynthesis.Figure 3.4 (a) Side and top view of the light‐harvesting complex LH2 from the purple bacterium R. acidophila 10050 (PDB: 1KZU [24]), showing bacteriochlorophyll a pigments in green and carotenoids (rhodopin glucoside) in orange. (b) Rod structure of c‐phycocyanin from the phycobilisome light‐harvesting antenna of cyanobacterium T. vulcanus (PDB: 3O18 [25]). (c) Light‐harvesting complex II (LHCII) from pea (PDB: 2BHW [26]), showing Chl a pigments in dark green and Chl b in light green.Figure 3.5 Examples of synthetic approaches to multi‐chromophore arrays: (a) a nine‐porphyrin array unit comprising a central free‐base porphyrin core that acts as final acceptor and is surrounded by eight energy‐donating zinc porphyrins [43]. Source: Choi et al. [41]Figure 3.6 Simplified Z‐scheme of natural oxygenic photosynthesis, showing how two photons are used per electron flowing from the terminal donor (H2O) to the terminal acceptor (NADP+) of the light‐dependent reactions.Figure 3.7 Photosystem II from T. vulcanus (PDB ID: 3WU2, a) and major redox‐active components within a PS‐II monomer involved in the main‐pathway electron transfer indicated with red arrows (b).Figure 3.8 The cycle of intermediate oxidation states of the oxygen‐evolving complex. Water is assumed to bind at the S3 state of the cluster and upon reconstitution of the S0 state. The S4 state is a postulated but unobserved transient intermediate that decays spontaneously to S0 with release of dioxygen.Figure 3.9 The Mn4CaO5 cluster and its protein pocket in the dark‐stable S1 state as revealed by protein crystallography (PDB ID: 3WU2, a), and a scheme showing the commonly used labeling of the ions comprising the inorganic core.Figure 3.10 Proposed models for the inorganic core in the S2 state of the OEC, with the first coordination sphere mostly omitted for clarity. The different magnetic topologies of the two valence isomers, as expressed through the pairwise exchange coupling constants Jij (values shown in cm−1; J < 0 is antiferromagnetic coupling), lead to different total spin states and g values for the corresponding EPR signals.Figure 3.11 The S2 → S3 transition according to Retegan et al. [191] and possible isovalent Mn(IV)4 components of the S3 state; the superscript “W” indicates binding of an additional water ligand.Figure 3.12 Two selected scenarios for the nature of the S4 state and OO bond formation from the computational literature: (a) formation of a Mn1(IV)‐oxyl group in the S4 state is followed by odd‐electron radical oxyl–oxo coupling [285], and (b) formation of a five‐coordinate high‐spin Mn4(V)‐oxo is followed by intramolecular nucleophilic coupling with concerted water binding [289]. Thick lines indicate direction of Jahn–Teller axes of Mn(III) ions.Figure 3.13 Reaction mechanism of RuBisCO proposed by Taylor and Andersson. Source: Taylor and Andersson [351].

3 Chapter 4Figure 4.1 Schematic illustration of photocatalytic hydrogen generation over a semiconductor photocatalyst [8].Figure 4.2 Band levels of various semiconductor photocatalysts. Source: Lu et al. [15].Figure 4.3 Schematic cells for rutile phase (a) and anatase phase (b) of TiO2. Source: Thompson and Yates [16].Figure 4.4 Photocatalytic H2 evolution of pure TiO2, Degussa P25, and TiO2 with different amounts of Zn. Source: Al‐Mayman et al. [25].Figure 4.5 Hydrogen evolution on the bare and metallic TiO2 photocatalysts using the benchmark Degussa P25 TiO2 and its bimetallic materials as the references. Source: Wang et al. [26].Figure 4.6 Photocatalytic hydrogen evolution on TiO2‐supported Au–Pd nanoparticles using a range of alcohols. Source: Su et al. [27].Figure 4.7 Schematic representation of the band structure of pure and N‐doped anatase TiO2. Note that the energies are not in scale. Source: Di Valentin et al. [32].Figure 4.8 The photocatalyst H2 generation rate of 2.5 wt%‐Cu2O/TiO2 with different sacrificial reagents. Source: Li et al. [34].Figure 4.9 (a) The energy band structure of ZnO/ZnS heterojunction. (b) The graphic structure of ZnO‐dotted ZnS. Source: Wu et al. [37].Figure 4.10 Tri‐s‐triazine‐based structure of g‐C3N4. The C and N atoms are indicated by gray and blue balls [8].Figure 4.11 Photocatalytic H2 production over the bulk C3N4, C3N4 NSs, and C3N4 NTs under visible‐light irradiation. Source: Zhu et al. [54].Figure 4.12 Graphic design of preparation of cobalt‐doped g‐C3N4. Source: Chen et al. [57].Figure 4.13 (a) EIS, (b) photocurrent response, and steady‐state (c) and transient (d) PL spectra of g‐C3N4 and B‐doped g‐C3N4 samples. Source: Chen et al. [60].Figure 4.14 Schematic illustration of the charge transfer for the three types of heterojunctions [8].Figure 4.15 Graphic illustration of the proposed mechanism for cobalt oxide/C3N4 NT heterojunction for photocatalytic H2 evolution. Source: Zhu et al. [61].Figure 4.16 Schematic diagram of the photocatalytic mechanism in WO3/MCN. Source: Kailasam et al. [63].Figure 4.17 The average H2 evolution rate of the SiC nanowires and modified SiC nanowires. Source: Hao et al. [66].Figure 4.18 Illustration of lanthanide upconversion nanoparticles (UCNPs).Figure 4.19 Illustrative diagrams of energy transfer among NYFG(15)/C3N4 NTs. Source: Zhu et al. [67].Figure 4.20 The proposed mechanism of the improved photocatalytic activity in the N‐CDs/CdS photocatalyst. Source: Shi et al. [69].Figure 4.21 Graphic diagram of the FeS2–TiO2 heterostructures under ultraviolet, visible, and NIR light irradiation for photocatalytic H2 evolution. Source: Kuo et al. [71].Figure 4.22 (a) The UV–vis near‐infrared absorption spectrum. (b) Photocatalytic hydrogen evolution for amorphous TiO2–x. Source: Jiang et al. [74]Figure 4.23 Methanol and its dissociative species underwent two‐electron oxidation in photocatalytic H2 production process over Au–Pt alloyed TiO2 nanocomposites. Source: Al‐Mayman et al. [25].

4 Chapter 5Figure 5.1 The schematic illustration of the features of H2 production by PV + EC, PEC, and PC in the five aspects of cost, STH efficiency, technology readiness, H2/O2 separation, and catalyst stability.Figure 5.2 The schematic illustration of three steps in PEC water splitting process. Source: From Wang and Wang [7]. © 2018 Elsevier.Figure 5.3 The illustration of semiconductor–electrolyte interface and the corresponding band bending. (a, b) The scheme and energy level of semiconductor and electrolyte in vacuum. The electrons are uniformly distributed around the core of atoms (blue dot). (c, d) The schematic illustration of electron distribution and the energy band bending when SEI is built. The electrons (red dot) are accumulated at the SEI interface. (e, f) The schematic illustration of electron–hole separation and transfer under illumination.Figure 5.4 (a) The oxygen vacancy concentration change by adjusting the treatment duration in N2. (b) The volcano relationship between photocurrent and oxygen vacancy concentration. Source: Wang et al. [39]. © 2019 Willey.Figure 5.5 (a) The schematic charge transfer in particulate Ta3N5 photoanode. (b) The change of charge separation and transfer efficiency for Ta3N5 photoanode when the charge transfer and generation are improved gradually. Source: Wang et al. [45]. Licensed under CC BY 3.0 Unported.Figure 5.6 (a) The image and (b) the photoresponse (red curve) of branched 2D porous TiO2 single‐crystal nanosheet photoelectrode. Insert (a) illustrates the photocharge transfer during PEC process. Source: Butburee et al. [71]. Reproduced with the permission from Wiley.Figure 5.7 (a) The surface state distribution characterization by cyclic voltammetry (CV) and the surface state related capacitance (Css). The different hematite photoelectrodes are compared including Ti doping, Al2O3 passivation, and H2O2 treatment. (b) The schematic illustration of PEC reaction via the surface states as intermediates. Source: From Wang et al. [8]. © 2016 Royal Society of Chemistry.Figure 5.8 The schematic illustration of cocatalyst design strategy based on a charge storage layer between the semiconductor and the effective cocatalyst layer.Figure 5.9 The schematic illustration of three different types of unbiased PEC water splitting systems. PA, PC, and PV represent photoanode, photocathode, and photovoltaic, respectively. The jE curves represent their photoresponses in a three‐electrode system. The intersections (red dots) indicate the unbiased operation points.

5 Chapter 6Figure 6.1 Schematic illustration of photocatalytic oxygen evolution systems: (a) homogeneous and (b) heterogeneous configuration.Figure 6.2 Structures of (a) PS‐II‐OEC natural WOCs and (b) λ‐MnO2, (c) Mn4O4L6 core, and (d) Co4O4(Ac)4(py)4 artificial WOCs. Source: From McCool et al. [27]. © 2011 American Chemical Society.Figure 6.3 Ball‐and‐stick model of (a) Mn4POM core, (b) polyhedral Mn4POM, and (c) S0 state of the natural OEC model. (d) Cycle diagram for the electron transfer within the S0 → S4 Kok cycle of the natural PS‐II‐OEC. (e) Photocatalytic oxygen evolution of Mn4POM within the [Ru(bpy)3]2+/Na2S2O8 system. Source: From Al‐Oweini et al. [20]. © 2014 John Wiley & Sons.Figure 6.4 Illustration of the band positions of common semiconductor materials and the energy level toward water splitting.Figure 6.5 Comparison of electronic DOS of (a) 3D bulk semiconductor materials and LD semiconductor materials including (b) 2D, (c) 1D, and (d) 0D. The inset arrows on related models indicate the quantum confinement direction and dimensionalities. (e) Quantum size effect on band structure. (f) Schematic illustration of surface plasmon resonance excitation on metallic NPs.Figure 6.6 (a) TEM image of WO2.83 NRs. (b) LSPR in 1D WO2.83 NRs (experimental measurement and theoretical simulation). Source: Manthiram and Paul Alivisatos [55].Figure 6.7 (a–c) TEM images and (d) size distributions of Co3O4 NPs with different diameters (3, 10, and 40 nm). (e) The relationship between particle diameter and surface area. Inset: Images of NPs in the form of powder or dispersed in water. (f) Visible‐light‐ driven oxygen evolution of Co3O4 NPs with different diameters using Ru(bpy)32+ as the sensitization system. (g) Visible‐light‐driven oxygen evolution from various cobalt compounds placed on SBA‐15. Source: Reproduced with permission. Grzelczak et al. [69]. Copyright 2013, American Chemical Society.Figure 6.8 (a) TEM image and (b) Mott–Schottky plots for CoO NPs and micropowders. The flat band potentials are obtained via extrapolation method. (c) Band positions of CoO NPs and micropowders according to the band gaps and flat band potentials. Source: Reproduced with permission Liao et al. [77]. Copyright 2014, Nature Publishing Group.Figure 6.9 (a) TEM image quantum‐sized BiVO4. (b) Comparison of the band position of nanoscale BiVO4 and quantum‐sized BiVO4. Source: Reproduced with permission. Sun et al. [89]. Copyright 2014, American Chemical Society.Figure 6.10 (a) Tungsten oxide single‐crystal nanosheets. (b,c) HRTEM images of WO3, WO3−x‐VT, and WO3−x‐HT nanosheets. (d) Band position regulation of WO3 nanosheets with oxygen vacancies. Source: From Yan et al. [58]. © 2015 John Wiley & Sons.Figure 6.11 (a) Representation of charge transfer between (001) and (110) facets. (b) Detachment and transfer of photoexcited charges in the BiOCl and defect‐rich BiOCl. (c) TEM image of the defect‐rich BiOCl ultrathin nanosheets. (d) Photocatalytic OER for BiOCl and defect‐rich BiOCl. Source: Reproduced with permission. Di et al. [107]. Copyright 2017, the Royal Society of Chemistry.Figure 6.12 The photocatalytic properties of LD materials including light absorption and charge transfer can be improved by doping engineering, active facet exposure, defect or vacancy creation, and LSPR. Hybrid LD materials with heterojunction can also be constructed in the form of 0D/1D, 1D/1D, 0D/2D, and 2D/2D configurations.Figure 6.13 (a,b) TEM images. (c) Time dependence of the O2 evolution. (d) Proposed energy diagram for Co(OH)2/TiO2. Source: Reproduced with permission. Maeda et al. [110]. Copyright 2016, Wiley‐VCH.

6 Chapter 7Figure 7.1 PEC water splitting by n‐type TiO2 photoanode for oxygen evolution reaction (OER) and Pt cathode for hydrogen evolution reaction (HER). The TiO2 with band‐gap energy (Eg) of 3.0 eV is photoexcited under ultraviolet light irradiation. External voltage (ΔEapp) is applied between the electrodes to induce water photoelectrolysis.Figure 7.2 Current–potential curves for conventional water electrolysis in the dark. Over potentials (η) are required for HER and OER over each electrocatalyst electrode (ηHER + ηOER).Figure 7.3 Current–potential curves for water photoelectrolysis using photoanodes for OER and a cathode for HER at the current density (j) of 2 mA cm−2. (a) Unbiased PEC water splitting and (b) PEC water splitting with externally applied voltage (ΔEapp). The ΔEphoto is the shift of the potential by using photoanodes in comparison with (c) anode electrocatalyst for OER. The absolute value of the current is equal between the series‐connected electrodes.Figure 7.4 (a) Current–potential curves of WO3 photoanode in 0.1 mol l−1 Na2SO4 (pH = 7) with 10 vol% methanol under UV irradiation (monochromatic light, λ = 365 nm) with different intensities of incident light. The light intensity (I0) was controlled by using a neutral density filter. (b) Effect of the I0 on the photocurrent density (jphoto) of the WO3 photoanode in the linear sweep voltammetry. The jphoto was linearly increased with I0 at each applied potential.Figure 7.5 (a) Flat band state and (b) band bending state of n‐type semiconductor. The recombination of the photoexcited e and h+ pair easily occurs in the flat band state. In contrast, the e and h+ pair is efficiently separated in the space charge layer (SCL). In the absence of surface states, the potential drop in the SCL (Δφ) is linear to the anodic shift of the applied potential from the flat band potential (Efb).Figure 7.6 The pH dependence of the electrode potentials at 25 °C. The potentials of SHE and Ag/AgCl reference electrode are constant, but the potentials of RHE and the potentials for HER and OER are linearly increased by an increase in pH value with a slope of 59 mV (ln 10 × RT/F) at 25 °C.Figure 7.7 IPCE action spectra of rutile and anatase TiO2 electrodes in 0.2 mol l−1 Na2SO4 with phosphate buffer (pH = 7) at 0.9 V vs. RHE. The inset shows their diffuse reflectance UV–vis spectra using Kubelka–Munk function, F(R). Source: From Amano et al. [21]. © 2018 The Electrochemical Society.Figure 7.8 Mott–Schottky plots of WO3 electrode in 0.1 mol l−1 H2SO4 (pH = 1) in the dark to estimate the flat band potential (Efb) and the donor density (ND). The capacitance of the space charge layer (Csc) was measured at 1 kHz with a sinusoidal amplitude of 10 mV in the dark. The inset shows the SEM image of the WO3 electrode surface. Source: Adapted with permission Amano et al. [22]. Copyright 2011, Springer‐Verlag.Figure 7.9 Linear sweep voltammograms of the WO3 electrode for (a) water oxidation in 0.1 mol l−1 H2SO4 and (b) methanol oxidation in the solution with 10 vol% methanol under photoirradiation (solid jE curves) and in the dark (dashed jE curves). The applied potential where photocurrent response starts is denoted as Eonset. The Efb is the value measured in Figure 7.8. Source: Based on Amano et al. [22].Figure 7.10 Band diagrams of n‐type semiconductors for PEC oxygen evolution: SrTiO3, rutile TiO2, monoclinic WO3, α‐Fe2O3 (hematite), monoclinic scheelite‐type BiVO4 (clinobisvanite) [28], and TaON [29] with E°(H+/H2) and E°(O2/H2O). Here, the reported Efb is considered as the conduction band (CB) minimum, although the CB minimum is more negative than Efb by 0.1–0.3 eV. The valence band (VB) maximum is determined from the CB minimum and the optical Eg.Figure 7.11 Pourbaix diagram, which is a potential–pH equilibrium diagram, for (a) tungsten–water system and (b) iron–water system at 25 °C. The molar concentration of WO42−, Fe2+, and Fe3+ is 10−6 mol l−1. The dashed lines show the potentials of HER and OER relative to pH value in water. Source: From Pourbaix [31]. © 1974, NACE.Figure 7.12 Schematic illustrations of photoexcited electron transport in photoanodes composed of (a) nanocrystalline particles with grain boundaries and (b) single‐crystalline materials with anisotropic nanostructures. Source: From Amano et al. [35]. © 2011 The Electrochemical Society.Figure 7.13 Crystal structure of WO3·H2O and cross‐sectional sideview SEM image of vertically aligned WO3 flakes deposited on transparent conductive oxide (TCO) glass substrate for PEC water oxidation under visible‐light irradiation. The seed layer, which is WO3 nanoparticles, is essential for heterogeneous nucleation of WO3·H2O flakes. The WO3·H2O flakes with two‐dimensional nanostructure were converted to monoclinic WO3 by heat treatment. Source: From Amano et al. [36]. © 2010, RCS.Figure 7.14 Current–potential curves of (a) vertically aligned WO3 flakes (blue dashed) and (b) horizontally stacked WO3 flakes (black solid) deposited on TCO glass substrate in 0.1 mol l−1 Na2SO4 (pH = 7) under visible‐light irradiation (λ > 400 nm). (A) Back‐side and (B) front‐side illumination were performed through the TCO glass substrate and the WO3 layer, respectively. Chopped illumination was used to show the transient photocurrent response for water oxidation. Source: Based on Amano et al. [35].Figure 7.15 (a) Pictures of the thermally oxidized TiO2 films (Ti plate calcined in air at 900 °C for two hours) and the films treated in H2 stream at different temperatures (300–800 °C). (b) Effect of H2 treatment temperature on the sheet resistance (R) and the donor density (ND) of TiO2 films. The R was measured by a four‐point probe. The ND was measured by the Mott–Schottky analysis. (c) Effect of H2 treatment temperature on the jphoto of TiO2 films in 0.1 mol l−1 H2SO4 (pH = 1). The jphoto for water oxidation was obtained in LSV measurement at 0.87 V vs. RHE under UV irradiation (λ > 300 nm). Source: Adapted with permission Amano et al. [38]. Copyright 2016, American Chemical Society.Figure 7.16 Schematic illustration of the PEC water oxidation process at the photoanode with cocatalyst layer: (a) conductive substrate, (b) n‐type semiconductor as a photoabsorber, (c) cocatalyst layer for OER, and (d) aqueous electrolyte solution. The number indicates the order of the PEC reaction process: (i) the generation of photoexcited eh+ pairs; (ii) charge separation, carrier diffusion, and carrier transport; and (iii) h+ transfer from semiconductor to water.Figure 7.17 Band alignments in three types of semiconductor heterojunctions: (a) straddling gap (type I), (b) staggered gap (type II), and (c) broken gap (type III).Figure 7.18 Schematic illustrations of (a) PEC cell for solar H2 production using water vapor from the air over the sea and (b) vapor‐fed PEC water splitting system using a gas‐diffusion HER cathode, a proton exchange membrane (PEM), and a gas‐diffusion photoanode for OER. The photoanode is composed of a TiO2 nanotube array decorated on porous Ti felt. The surface of TiO2 nanotubes is coated with Nafion ionomer thin film for the gas‐phase operation. Source: (a) Adapted by permission of Wiley‐VCH Verlag. Amano et al. [53]; (b) Based on Amano et al. [54].Figure 7.19 Vapor‐fed water photoelectrolysis by porous SrTiO3 photoanode |PEM| Pt‐carbon black cathode under humidified argon (3 vol% water vapor). The response of (a) photocurrent density and (b) the formation rate of H2 evolved in the cathode compartment and O2 evolved in the photoanode compartment at ΔEapp = 0.3 V under 365‐nm UV irradiation (I0 = 42 mW cm−2, photoanode area 2 cm2). The porous SrTiO3 photoanode was coated by Nafion ionomer thin films for the gas‐phase operation. The evolved gas in each compartment was separated by a PEM and analyzed by on‐line gas chromatographs. Source: (a, b) Based on Amano et al. [54].

7 Chapter 8Figure 8.1 Different types of photocatalytic overall water splitting systems including (a) one‐step photoexcitation, (b) first‐generation Z‐scheme, (c) second‐generation Z‐scheme, and (d) third‐generation Z‐scheme. HEP: hydrogen evolution photocatalyst; OEP: oxygen evolution photocatalyst; HEC: hydrogen evolution cocatalyst; OEC: oxygen evolution cocatalyst; CB: conduction band; VB: valence band; Eg: energy band gap; NHE: normal hydrogen electrode; Ox: oxidant; Red: reductant.Figure 8.2 Representative material types for one‐step photocatalytic overall water splitting systems. (A) SrTiO3 with 18 facets and selective deposition of HER and OER cocatalysts. Source: Mu et al. [68]. Copyright 2016, the Royal Society of Chemistry.Figure 8.3 Representative Z‐scheme types for photocatalytic overall water splitting. (a) First‐generation Z‐scheme coupling ZrO2/TaON(HEP) and BiVO4(OEP) with Fe3+/Fe2+ redox pairs as aqueous electron mediator and corresponding performance for 12 hours. Source: From Qi et al. [237]. © 2018 Elsevier.Figure 8.4 Different types of PEC for water splitting comprising (a) photoanode–cathode, (b) photocathode–anode, (c) photoanode–photocathode tandem cell, and (d) photovoltaic–photoanode (PV–PEC). Source: From Kim et al. [19]. © 2019 RCS.Figure 8.5 Photoanode–photocathode tandem cell for overall water splitting systems using (a) metal oxide‐based photoelectrodes (Cu2O as HEP and Mo:BiVO4 as OEP) and produced photocurrent density with stoichiometric H2 and O2 evolution. Source: Reproduced with permission. Pan et al. [368]. Copyright 2018, Macmillan Publishers Limited, part of Springer Nature.Figure 8.6 Photoanode–photocathode tandem cell for overall water splitting systems using (A) PEC system composed of IrOx/n‐GaAs (photoanode) and Pt/Ti/Pt/Au/p‐GaAS (photocathode) and representative JE curve. Source: Reproduced with permission. Kang et al. [395]. Copyright 2017, Macmillan Publishers Limited, part of Springer Nature.Figure 8.7 Representative PV–PEC devices for overall water splitting. (a) Inverted metamorphic multi‐junction (IMM) device structure with GaInP/GaInAs tandem PV cells coupled with RuOx–Pt and J–V measurements. Source: From Young et al. [403]. © 2017 Springer Nature.Figure 8.8 Representative PV–PEC devices for overall water splitting. (a) System composed of single‐junction perovskite solar cell and a nanocone/Mo‐doped BiVO4/Fe(Ni)OOH photoanode, and respective J–V curve for the tandem device. Source: From Qiu et al. [325]. © 2016 Yongcai Qiu.Figure 8.9 Representative artificial (wireless) PV–PEC devices for overall water splitting. (a) Tandem CH3NH3PbI3 perovskite single‐junction solar cell with Co‐Ci incorporated H‐doped Mo/BiVO4 photoanode and corresponding gas evolution with calculated STH efficiency. Source: From Kim et al. [421]. © 2015 American Chemical Society.

8 Chapter 9Figure 9.1 Schematic illustration of the natural photosynthesis (a) and the artificial photosynthesis (b) using semiconductor as the photocatalyst.Figure 9.2 Band positions of some semiconductor photocatalysts and the redox potentials of CO2 reduction at pH 7 in aqueous solution. Source: Li et al. [12]. © 2014, Springer Nature.Figure 9.3 Schematic diagram of the slurry reactor (a) and the fixed bed reactor (b). Source: (a) Ola and Maroto‐Valer [17]. Licensed under CC BY 4.0. (b) Bakherad et al. [19]. © 2020, Royal Society of ChemistryFigure 9.4 (a) Schematic illustration of optical‐fiber photo reactor. (b) Light propagation in a TiO2‐coated fiber reactor. Source: Wu et al. [20]. © 2008, Springer Nature.Figure 9.5 Dye‐sensitized photocatalytic reaction.Figure 9.6 Band alignment of bulk CdSe, CdSe QDs with a diameter of 2.5 nm, and TiO2, as well as relevant redox potentials of CO2 and H2O. Source: Wang et al. [25]. © 2010, American Chemical Society.Figure 9.7 Schematic diagram of the reduction of CO2 and water vapor into hydrocarbon fuels through a nanotube–catalyst array. Source: Varghese et al. [28]. © 2009, American Chemical Society..Figure 9.8 Semiconductor heterojunctions in the form of type I (a), type II (b), and type III (c).Figure 9.9 Schematic diagram of the photoexcitation process of AgBr/TiO2 composite under visible‐light irradiation.Figure 9.10 Schematic illustration of TiO2 with {001} and {101} surface heterojunction. Source: Yu et al. [31]. © 2014, American Chemical Society.Figure 9.11 Schematic illustration of the spatial separation of redox sites on the TiO2 photocatalysts prepared without HF (a), by adding a moderate amount of HF (b), and by adding a high amount of HF (c).Figure 9.12 Schematic diagram of charge transfer in a Z‐scheme semiconductor–semiconductor composite.Figure 9.13 Schematic diagram of Z‐scheme BiVO4/C/Cu2O nanowire arrays. Source: Reproduced with permission. Kim et al. [32]. © 2018, American Chemical Society.Figure 9.14 Comparison of the energy band positions and the related redox reaction potentials of g‐C3N4 and ZnO. Source: Yu et al. [33]. © 2015, Royal Society of Chemistry.Figure 9.15 Schematic diagram of the solid–liquid interface structure.Figure 9.16 Potential comparison about the reduction of CO2, H2CO3, and CO32− ions into methanol. Source: Pan and Chen [38]. © 2007, Elsevier.Figure 9.17 Possible adsorption mechanism of CO2 on TiO2 surface. Source: Wu and Huang [40]. © 2010, Springer Nature.Figure 9.18 Schematic illustration of the charge transfer and separation in RGO–CdS nanorod system under visible‐light irradiation. Source: Yu et al. [41]. © 2014, Royal Society of Chemistry.Figure 9.19 Two proposed mechanisms for the photoreduction of CO2 to methane: formaldehyde (a) and carbine (b) pathways. Source: Izumi [42]. © 2015, American Chemical Society.Figure 9.20 Photocatalytic reduction of CO2 (86.65 kPa) by TiO2 (a) and Pd(1%)‐TiO2 (b). Source: Xiong et al. [44]. © 2017, Elsevier.Figure 9.21 Schematic mechanism of photocatalytic CO2 reduction over CQDs/OCN‐x under visible‐light irradiation. Source: Li et al. [45]. © 2019, Springer Nature.

9 Chapter 10Figure 10.1 Schematic diagrams for PEC CO2 reduction in water using a semiconductor as (a) photocathode, (b) photoanode, and (c) both photoanode and photocathode. (d) Schematic diagram for the device combining a photovoltaic cell with an efficient electrochemical catalyst for CO2 reduction. Source: Zhang et al. [18]. © 2018, Springer Nature.Figure 10.2 (a) Scanning electron microscope (SEM) image of HA‐Co3O4. (b) Upper layer microflower morphology. (c) Lower layer rhombus nanorod morphology. (d) Transmission electron microscopy (TEM) of petals on the flower; Inset: selected area electron diffraction (SAED) pattern of petal region. (e) High‐resolution transmission electron microscopy (HRTEM) of petal region. (f) Formate yields of HA‐Co3O4 under PEC and EC under different negative potentials. Crystallographic modeling of Co3O4 (g) {12}, (h) {001}, (i) {110}, and (j) {111} facets with side view (left) and oblique view (right). Blue, light blue, and red spheres represent Co3+, Co2+, and O2−, respectively. Source: Huang et al. [54]. © 2013, American Chemical Society.Figure 10.3 (a) Linear sweep voltammetry of the BCN3.0 electrodes loaded with or not loaded with cocatalysts. (b) Products analyses of photoelectrochemical reduction of CO2 over cocatalyst‐loaded BCN3.0 electrodes. Source: Sagara et al. [26]. © 2016, Elsevier.Figure 10.4 Periodic table depicting the primary reduction products in CO2‐saturated aqueous electrolytes on metal and carbon electrodes. Source: White et al. [2]. © 2015, American Chemical Society.Figure 10.5 Schematic representations for (a) CoII complex and (b) FeII complex for PEC reduction of CO2. Source: Chen et al. [83]. © 2015, American Chemical Society.Figure 10.6 Schematic of (a) the ruthenium dye‐sensitized TiO2 nanoparticles with adsorbed CODH enzyme, which catalyzes the reduction of CO2 to CO. Source: Woolerton et al. [89]. © 2010, American Chemical SocietyFigure 10.7 Schematic image of (a) NiO–RuRe as photocathode and CoOx/TaON as photoanode. Source: Sahara et al. [30]. © 2016, American Chemical Society.

10 Chapter 11Figure 11.1 (a) Schematic illustrations of different pathways for N2 reduction, including the dissociative pathway, associative alternating pathway, and associative distal pathway. Source: (a) Shipman et al. [13]. © 2017, Elsevier.Figure 11.2 (a) Schematic of the plasmon excitation‐based photocatalytic N2 reduction. (b) TEM image of the Au/TiO2−OV NSs. (c) EPR spectra of four catalysts of TiO2, TiO2−OV, Au/TiO2, and Au/TiO2−OV. (d) Detected NH3 concentrations of four catalysts. Source: (a–d) Adapted from Yang et al. [30]. © 2018, American Chemical Society.Figure 11.3 (a) Schematic of the synthesis process of engineering NVs into g‐C3N4. (b) ESR and (c) TPD spectra of the prepared g‐C3N4 and V−g‐C3N4. Source: (a–c) Adapted from Dong et al. [43]. © 2015, Royal Society of Chemistry.Figure 11.4 (a) Schematic photocatalytic N2 reduction on ultrathin MWO‐1 nanowires. Adsorption configurations of N2 molecules with different charge dispersions on (b) defect‐rich W18O49, and (c) Mo‐doped W18O49. (d) O K‐edge XAS spectra of different photocatalysts. (e) Schematic of the PCET process for photocatalytic N2 fixation. (f) Photocatalytic NH3 yield rates with MWO-1. Source: (a–f) Adapted from Zhang et al. [53]. © 2018, American Chemical Society.Figure 11.5 (a) and (b) HAADF−STEM images and (c) EDX mapping of A‐SmOCl. (d) ESR signals and (e) O K‐edge XAS spectra of different photocatalysts. (f) In situ DRIFT spectra recorded in the N2 reduction process. Source: (a–f) Hou et al. [61]. © 2019, Elsevier.Figure 11.6 (a) SEM and (b) TEM images of the GaN NWs. (c) TEM image of Ru (5 wt%) decorated GaN NWs. Inset: the diameter distributions of Ru nanoclusters. (d) Schematic of the Schottky barrier between Ru clusters and n‐type GaN NWs. (e) Schematic illustration for the InGaN/GaN with five segments of InGaN on the template of GaN nanowire. (f) NH3 evolution rates of various photocatalysts under visible‐light illumination. Source: (a–f) Li et al. [74]. © 2017, John Wiley & Sons.Figure 11.7 (a) Schematic illustration of photocatalytic cell for N2 fixation. (b) and (c) HRTEM images of hollow Au–Ag2O, the inset is the corresponding FFT of the nanoparticle. (d) NH3 yields and solar‐to‐ammonia (STA) efficiencies of different photocatalysts. (e) NH3 yields of Au–Ag2O under various operating conditions. Source: (a–e) Nazemi et al. [82]. © 2019, Elsevier.Figure 11.8 (a) Schematic of N2 reduction on the aerophilic–hydrophilic interface. (b) Droplet shapes of the liquid and N2 bubble on different interfaces. (c) Interfacial water molecule spectra on the surface of Au-PTFE/TS . (d) NH3 yield rates and FEs on Au/TS and Au−PTFE/TS at different potentials. Source: (a–d) Zheng et al. [89]. © 2019, Elsevier.Figure 11.9 (a) Schematic of the synthesis steps for the Au/end−CeO2 heterojunction. (b) HAADF−STEM image and the related elemental mapping. Mechanism for N2 photo‐fixation on the Au/end−CeO2 heterojunction. The hot carrier separation behaviors on (c) the Au/end−CeO2 and (d) the core@shell nanostructures. Source: (a–d) Jia et al. [111]. © 2019, American Chemical Society.Figure 11.10 (a) Schematic of TiO2−xHy/Fe catalyst for dual‐temperature‐zone photo‐thermal NH3 synthesis. (b) High‐resolution STEM image of TiO2−xHy/Fe catalyst. (c) Steady‐state non‐equilibrium temperature distribution and dual‐temperature‐zone NH3 synthesis on TiO2−xHy/Fe. (d) LTDs between hot Fe and cooling TiO2−xHy. (e) SERS mapping of the spatial dispersion of TiO2−xHy, Fe, and the adonitol and the detected local LTD. (f) The NH3 concentrations produced at different apparent catalyst temperatures. (g) Successive light‐on/off measurement of photo‐thermal NH3 synthesis. (h) The impressive equilibrium‐beyond reactivity of TiO2−xHy/Fe at different pressures. Source: (a–h) Mao et al. [118]. © 2019, Elsevier.Figure 11.11 (a) HRTEM, (b) elemental mapping, and (c) HAADF−STEM images of Mo−PCN. (d) Time‐resolved fluorescence kinetics. (e) LT−FTIR spectra for N2 adsorption. (f) Charge density difference of Mo−PCN with adsorbed N2. Source: (a–f) Guo et al. [119]. © 2019, Royal Society of Chemistry.Figure 11.12 (a) Schemcatic of designed Mo2Fe6S8−Sn2S6 biomimetic chalcogel. Source: (a) Adapted from Banerjee et al. [122]. © 2015, American Chemical Society.

11 Chapter 12Figure 12.1 (a) UV–vis DRS and (b) N2 adsorption isotherms at 77 K for MIL‐125‐NH2 and NiO/MIL‐125‐NH2. Source: Isaka et al. [24]. © 2018, Royal Society of Chemistry.Figure 12.2 Ni K‐edge (a) XAFS and (b) FT‐EXAFS spectra of NiO/MIL‐125‐NH2. Source: Isaka et al. [24]. © 2018, Royal Society of Chemistry.Figure 12.3 (a) TEM image, (b) HAADF‐STEM image, and (c‐f) EDX mappings of NiO/MIL‐125‐NH2. Source: Reproduced with permission from Isaka et al. [24].Figure 12.4 (a) Time course of photocatalytic H2O2 production catalyzed by MIL‐125‐NH2 in the presence and absence of TEOA and visible‐light irradiation (λ > 420 nm). (b) Time courses of photocatalytic H2O2 production catalyzed by MIL‐125‐NH2, NiO/MIL‐125‐NH2, and Pt/MIL‐125‐NH2. Source: (a) Isaka et al. [24]. © 2018, Royal Society of Chemistry.Figure 12.5 (a) Time course of TEOA oxidation product formation under visible‐light (λ > 420 nm) irradiation for MIL‐125‐NH2. (b) Mass spectrum of the peak after photoirradiation. (c,d) Reported oxidation products of TEOA in literatures. (e) Plausible product of TEOA oxidation. Source: Isaka et al. [24]. © 2018, Royal Society of Chemistry.Figure 12.6 Time courses of H2O2 and benzaldehyde production of (a) MIL‐125‐NH2 and (b) NiO/MIL‐125‐NH2 dispersed in an acetonitrile solution of benzyl alcohol under visible‐light irradiation (λ > 420 nm). Source: Isaka et al. [24]. © 2018, Royal Society of Chemistry.Figure 12.7 Time courses of H2O2 (15 mM) decomposition dissolved in 5.0 ml of acetonitrile suspension of 5.0 mg of MIL‐125‐NH2 and NiO/MIL‐125‐NH2 at 313 K. Source: Isaka et al. [24]. © 2018, Royal Society of Chemistry.Figure 12.8 EPR spectral of suspensions containing DMPO and (a) MIL‐125‐NH2 or (b) NiO/MIL‐125‐NH2 under visible‐light irradiation (λ > 420 nm). Source: (a) Isaka et al. [24]. © 2018, Royal Society of Chemistry.Figure 12.9 (a) Digital photographs of two‐phase systems composed of an aqueous phase and a benzyl alcohol phase containing MIL‐125‐NH2 (left) and MIL‐125‐Rn (n = 4 and 7, right). (b) Photocatalytic H2O2 production in the two‐phase system. Source: Isaka et al. [28]. © 2019, John Wiley & Sons.Figure 12.10 (a) XRD patterns and (b) UV–vis DRS for MIL‐125‐NH2, MIL‐125‐R4, and MIL‐125‐R7. Source: Isaka et al. [28]. © 2019, John Wiley & Sons.Figure 12.11 FE‐SEM images of (a,c) MIL‐125‐NH2 and (b,d) MIL‐125‐R7 (a,b) before and (c,d) after three hours of photoirradiation (λ > 420 nm) reaction in the two‐phase system composed of benzyl alcohol (5.0 ml) and water (2.0 ml). Source: Reproduced with permission from Isaka et al. [28].Figure 12.12 TEM images of (a,c) MIL‐125‐NH2 and (b,d) MIL‐125‐R7 (a,b) before and (c,d) after three hours of photoirradiation (λ > 420 nm) reaction in the two‐phase system composed of benzyl alcohol (5.0 ml) and water (2.0 ml). Source: Reproduced with permission from Isaka et al. [28].Figure 12.13 TG‐DTA measurement of (a,b) MIL‐125‐NH2, (c,d) MIL‐125‐R4, and (e,f) MIL‐125‐R7. Chemical structures of linkers in (g) MIL‐125‐NH2, alkylated linker in (h) MIL‐125‐R4, and (i) MIL‐125‐R7. Source: Isaka et al. [28]. © 2019, John Wiley & Sons.Figure 12.14 (a) Water adsorption isotherms at 298 K for MIL‐125‐NH2, MIL‐125‐R4, and MIL‐125‐R7. The water contact angles of (b) MIL‐125‐NH2, (c) MIL‐125‐R4, and (d) MIL‐125‐R7. Source: (a) Isaka et al. [28]. © 2019, John Wiley & Sons; (b–d) Reproduced with permission from Isaka et al. [28].Figure 12.15 (a) Digital photograph of MIL‐125‐NH2 (left) and MIL‐125‐R7 (right) dispersed into the two‐phase system; aqueous phase (2.0 ml) was observed on top of the benzyl alcohol phase (5.0 ml). (b) Time courses of H2O2 production under photoirradiation (λ > 420 nm) in the two‐phase system catalyzed by 5.0 mg of catalysts. Source: (a) Reproduced with permission from Isaka et al. [28]; (b) Isaka et al. [28]. © 2019, John Wiley & Sons.Figure 12.16 Time courses of benzaldehyde formation under photoirradiation (λ > 420 nm) of the two‐phase system composed of benzyl alcohol (5.0 ml) and water (2.0 ml) catalyzed by 5.0 mg of catalysts. Source: Isaka et al. [28]. © 2019, John Wiley & Sons.Figure 12.17 XRD patterns of MIL‐125‐NH2 and MIL‐125‐R7 after three hours of photoirradiation (λ > 420 nm) reaction in the two‐phase system composed of benzyl alcohol (5.0 ml) and water (2.0 ml). Source: Isaka et al. [28]. © 2019, John Wiley & Sons.Figure 12.18 Recycling tests of H2O2 production under photoirradiation (λ > 420 nm) of the two‐phase system composed of benzyl alcohol (10.0 ml) and water (4.0 ml) catalyzed by 10.0 mg of MIL‐125‐NH2 (blue) and MIL‐125‐R7 (orange). Source: Isaka et al. [28]. © 2019, John Wiley & Sons.Figure 12.19 (a) Time courses of H2O2 production under photoirradiation (λ > 420 nm) of the two‐phase system composed of benzyl alcohol (5.0 ml) and water (2.0, 5.0, and 10.0 ml) catalyzed by 5.0 mg of MIL‐125‐R7. (b) Time courses of H2O2 production under photoirradiation (λ > 420 nm) of the two‐phase system composed of benzyl alcohol (5.0 ml) and water (2.0 ml) at different pH values (0.3, 1.3, 2.1, and 6.7) catalyzed by 5.0 mg of MIL‐125‐R7. (c) Time courses of H2O2 production under photoirradiation (λ > 420 nm) of the two‐phase system composed of benzyl alcohol (5.0 ml) and an aqueous phase (2.0 ml, deionized water or saturated NaCl aqueous solution) catalyzed by 5.0 mg MIL‐125‐R7. Source: Isaka et al. [28]. © 2019, John Wiley & Sons.Figure 12.20 N2 adsorption isotherms at 77 K for MIL‐125‐NH2, MIL‐125‐R4, and MIL‐125‐R7. Source: Isaka et al. [28]. © 2019, John Wiley & Sons.Figure 12.21 Photocatalytic H2O2 production with (a) linker‐alkylated MOF: MIL‐125‐R7 and (b) cluster‐alkylated MOF: OPA/MIL‐125‐NH2. Source: Kawase et al. [31]. © 2019, Royal Society of Chemistry.Figure 12.22 (a) XRD patterns and (b) UV–vis DRS of MIL‐125‐NH2 and OPA/MIL‐125‐NH2. Source: Kawase et al. [31]. © 2019, Royal Society of Chemistry.Figure 12.23 N2 adsorption isotherms at 77 K for MIL‐125‐NH2 and OPA/MIL‐125‐NH2. Source: Kawase et al. [31]. © 2019, Royal Society of Chemistry.Figure 12.24 (a) TG and (b) DTA profiles of OPA/MIL‐125‐NH2. (c) Chemical structures of alkylated clusters in OPA/MIL‐125‐NH2. Source: Kawase et al. [31]. © 2019, Royal Society of Chemistry.Figure 12.25 FTIR spectra of (a) MIL‐125‐NH2, OPA/MIL‐125‐NH2, and (b) OPA. Source: Kawase et al. [31]. © 2019, Royal Society of Chemistry.Figure 12.26 (a) XPS spectra of Ti 2p in MIL‐125‐NH2 and OPA/MIL‐125‐NH2. (b) XPS spectra of P 2p in MIL‐125‐NH2 and OPA/MIL‐125‐NH2 before and after etching. Source: Kawase et al. [31]. © 2019, Royal Society of Chemistry.Figure 12.27 (a) Water contact angle of OPA/MIL‐125‐NH2. (b) Time courses of H2O2 production of OPA/MIL‐125‐NH2 under photoirradiation (λ > 420 nm) in the two‐phase system. Source: (a) Kawase et al. [31]. © 2019, Royal Society of Chemistry; (b) Reproduced with permission from Kawase et al. [31].Figure 12.28 Time courses of benzaldehyde production of OPA/MIL‐125‐NH2 under photoirradiation (λ > 420 nm) in the two‐phase system. Source: Kawase et al. [31]. © 2019, Royal Society of Chemistry.Figure 12.29 Time courses of H2O2 production under photoirradiation (λ > 420 nm) of a single‐phase system composed of an acetonitrile solution (5.0 ml) of BA (1.0 ml) catalyzed by 5.0 mg of catalysts. Source: Kawase et al. [31]. © 2019, Royal Society of Chemistry.Figure 12.30 Recycling tests of MIL‐125‐NH2, MIL‐125‐R7, and OPA/MIL‐125‐NH2. Source: Kawase et al. [31]. © 2019, Royal Society of Chemistry.Figure 12.31 XRD patterns of OPA/MIL‐125‐NH2 before and after reaction. Source: Kawase et al. [31]. © 2019, Royal Society of Chemistry.Figure 12.32 N2 adsorption isotherms at 77 K for OPA/MIL‐125‐NH2 before and after the reaction. Source: Kawase et al. [31]. © 2019, Royal Society of Chemistry.

12 Chapter 13Figure 13.1 (a) The kinetic progress of methaneactivation on thermocatalysis and photocatalysis methods, (b) reaction mechanism of photocatalysis, and (c) reaction mechanism of photoelectrochemical catalysis. Source: Song et al. [2].Figure 13.2 (a–d) Photoexcitation and subsequent relaxation processes following the illumination of a metal nanoparticle with a laser pulse and characteristic time scales. Source: Brongersma et al. [28]Figure 13.3 (A) Schematic (a) a conventional vacuum line equipped with pressure gage; (b) joint; (c) small hole for thermo‐couple; (d) catalyst bed; (e) UV‐reflection mirror; (f) Xe lamp. drawing of the fixed bed photo‐reactor. Source: Yoshida et al. [46]Figure 13.4 (a) Difference on activation energy between photocatalysis and thermocatalysis. Source: Chen et al. [5], licensed under CC BY 4.0Figure 13.5 (a) Proposed mechanisms of photooxidation of CH4 on TiO2. Source: Li et al. [26] © 2018, American chemical society

13 Chapter 14Figure 14.1 Composition of biomass.Figure 14.2 (a) Reaction process images, quantification, and distribution of products obtained from the photocatalytic depolymerization of birch lignin. (b) Proposed mechanism for β‐O‐4 bond cleavage in the photocatalytic conversion of lignin over the Zn4In2S7 catalyst. Source: Reprinted with permission from Lin et al. [27].Figure 14.3 Photographs taken before (a) and after (b) precipitation of lignin during photoelectrocatalytic oxidation of 500 ppm lignin at different times, (c) Fourier transform infrared (FTIR) spectra of lignin and modified lignin, and (d) intermediates analyzed using HPLC. Source: Reprinted with permission from Tian et al. [48].Figure 14.4 Illustration of experimental reactor used for photoreforming cellulose to hydrogen via combined photocatalysis and acid hydrolysis. Source: Zou et al. [56].Figure 14.5 Photothermally promoted cleavage of β‐1,4‐glycosidic bonds of cellulose on Ir/HY catalyst. Source: Reprinted with permission from Zhang et al. [57].Figure 14.6 Schematic representation of the photocatalytic system for the H2 evolution by water splitting over irradiated Pt/TiO2 in the presence of cellulose as the sacrificial agent. Source: Speltini et al. [62], licensed under CC BY 3.0.Figure 14.7 Photocatalytic H2 production using activated NCNCNx (5 mg) and Ni bis(diphosphine) (NiP) (50 nmol) with purified lignocellulose components (100 mg) in potassium phosphate (KPi) solution (0.1 M, pH 4.5, 3 ml) under AM 1.5 G irradiation for 24 hours at 25 °C. Source: Kasap et al. [68].Figure 14.8 Schematic illustration of the lignocellulose structure and photocatalytic valorization of native lignin. Source: Reprinted with permission from Wu et al. [70].Figure 14.9 (a) Scanning electron microscope (SEM) image of sample Bi2WO6; (b) Time‐online photocatalytic performance toward selective oxidation of glycerol to DHA over Bi2WO6. Source: (a) Reprinted with permission from Zhang et al. [74]; (b) Zhang et al. [74].Figure 14.10 A simple solar‐induced hybrid direct glycerol fuel cell consists of a Pt cathode, a Na‐Pi buffer electrolyte with glycerol as fuel and NiPi/Pi–Fe2O3 as a photoanode, and the possible photo‐generated charging process of glycerol process of glycerol oxidation over NiPi/Pi–Fe2O3. Source: Chong et al. [97].Figure 14.11 SEM image of (a) W:BiVO4 and (b) pTTh films. (c) The energy level diagram of the dual‐photoelectrode PFC. Source: (a, b) Modified with permission from Zhang et al. [36]; (c) Zhang et al. [36].Figure 14.12 Schematic configuration and electron transfer processes of PEC glucose/glucose fuel cell employing TiO2/ITO photoanode and ITO cathode upon incorporation of UV light and MB. Insets show SEM images of TiO2/ITO and structures of MB and LMB. Source: Zhao et al. [104].Figure 14.13 A plausible reaction mechanism of PEC selective aerobic oxidation of benzyl alcohols. Source: Zhang et al. [112].Figure 14.14 A photoelectrochemical cell for HMF reduction. Inset is the SEM image showing the surface morphologies of Aggd. Source: (a) Roylance et al. [116]; (b) Reprinted with permission from Roylance et al. [116].

14 Chapter 15Figure 15.1 Schematic illustration of electron–hole pair generation in a semiconductor upon light irradiation. Possible pathways are labeled in 1– 4, where 1, reduction; 2, oxidation; 3, volume recombination; and 4, surface recombination; A, electron acceptor, and B, electron donor. Source: Linsebigler et al. [4].Figure 15.2 Schematic illustration of different photocatalytic water splitting systems. (a) One‐step photoexcitation or single‐component photocatalytic system. (b) Photocatalytic system with type‐II heterojunction. (c) Two‐step photoexcitation or Z‐scheme photocatalytic system. Source: Chen et al. [6].Figure 15.3 Possible configurations of CO2 absorption on the surface of photocatalysts. Source: Liu et al. [12].Figure 15.4 Bandgap energies and the band edge positions of some commonly reported photocatalysts as well as the products derived from photocatalytic CO2 reduction with reference to normal hydrogen electrode (NHE). Source: Tu et al. [10].Figure 15.5 (a) Photoanode–HEC cathode, (b) photocathode–OEC anode, (c) photoanode–photocathode tandem cell, and (d) photoanode–photovoltaic (PEC‐PV) tandem cell. Source: Kim et al. [1].Figure 15.6 Schematic illustration of side‐irradiated water splitting reactor. Source: Takata et al. [23].Figure 15.7 Schematic illustration of top‐irradiated water splitting reactor. Source: Kudo et al. [25].Figure 15.8 (a) Schematic illustrations of inner‐irradiated reactor in lab scale. Source: Wei et al. [27].Figure 15.9 Schematic illustration of a twin photo‐reactor system. Source: Sasaki et al. [29].Figure 15.10 Schematic illustration of photocatalytic water splitting panel. Source: (a) Reproduced with permission Goto et al. [30]. Copyright 2018, Elsevier; (b) Goto et al. [30].Figure 15.11 Schematic illustration of (a) slurry reactor, (b) internally illuminated reactor, and (c) optical fiber reactor. Source: Ola et al. [35], licensed under CC BY 4.0.Figure 15.12 Schematic diagram of photocatalytic CO2 reduction in a continuous gas flow fixed bed reactor. Source: Kong et al. [7].Figure 15.13 Schematic diagram of PEC water splitting system in the configuration of (a) n‐type photoanode‐HEC and (b) p‐type photocathode‐OEC. Source: Ahmed et al. [40].Figure 15.14 Schematic diagram of PEC water splitting devices: (a) with membrane and (b) without membrane. (c) Photograph of an integrated membrane‐free device with internal wiring. Source: Reproduced with permission Jin et al. [41]. Copyright 2014, Royal Society of Chemistry.Figure 15.15 Schematic diagram of PEC CO2 reduction system. Source: Cheng et al. [42].Figure 15.16 Large‐scale reactor for photocatalytic H2 production containing immobilized Pt/g‐C3N4 working under natural sunlight. Source: Reproduced with permission Schröder et al. [43]. Copyright 2015, Wiley.Figure 15.17 (a) Schematic illustration of synthesis protocol of SrTiO3:La,Rh/Au/BiVO4:Mo Z‐scheme photocatalyst sheets using particle transfer method. (b) Electron transfer mechanism of SrTiO3:La,Rh/Au/BiVO4:Mo Z‐scheme system. (c) Photograph of the ink used for screen printing and the corresponding printed photocatalyst sheets. Source: (a, c) Reproduced with permission Wang et al. [30]. Copyright 2016, Nature Publishing Group; (b) Wang et al. [30].Figure 15.18 (a) False‐colored SEM image of Si/TiO2 nanotree arrays photocatalyst sheets. (b) Photographs of Si/TiO2 photocatalyst sheets. (c) Charge transfer mechanism in Si/TiO2 photocatalyst sheets. Source: Reproduced with permission Liu et al. [44]. Copyright 2013, American Chemical Society.Figure 15.19 (a) Wired PEC cell and monolithic wireless cell of Co‐Bi/3jn a‐Si/NiMoZn system. Source: Reece et al. [45]. Copyright 2011, American Association for the Advancement of Science.Figure 15.20 (a) SrTiO3‐based artificial leaf with CuxO as HER co‐catalysts and CoPi as OER co‐catalysts. Source: Reproduced with permission Chen et al. [48]. Copyright 2017, American Chemical Society.

Solar-to-Chemical Conversion

Подняться наверх