Site-Selective Growth of Crystalline Ceria with Oxygen Vacancies on

Mar 22, 2019 - Site-selective growth of crystalline semiconductors on gold nanocrystals remains a .... The town and its surroundings enjoy a warm, mar...
0 downloads 0 Views 368KB Size
Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

Communication

Site-Selective Growth of Crystalline Ceria with Oxygen Vacancies on Gold Nanocrystals for NIR Nitrogen Photofixation Henglei Jia, Aoxuan Du, Han ZHANG, Jianhua Yang, Ruibin Jiang, Jianfang Wang, and Chun-yang Zhang J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.8b13062 • Publication Date (Web): 22 Mar 2019 Downloaded from http://pubs.acs.org on March 22, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 5 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Site-Selective Growth of Crystalline Ceria with Oxygen Vacancies on Gold Nanocrystals for NIR Nitrogen Photofixation Henglei Jia,† Aoxuan Du,† Han Zhang,‡ Jianhua Yang,‡ Ruibin Jiang,*,§ Jianfang Wang,*,‡ and Chunyang Zhang*,† College of Chemistry, Chemical Engineering and Materials Science, Collaborative Innovation Center of Functionalized Probes for Chemical Imaging in Universities of Shandong, Key Laboratory of Molecular and Nano Probes, Ministry of Education, Shandong Provincial Key Laboratory of Clean Production of Fine Chemicals, Shandong Normal University, Jinan 250014, China ‡ Department of Physics, The Chinese University of Hong Kong, Shatin, Hong Kong SAR, China § Shaanxi Engineering Lab for Advanced Energy Technology, School of Materials Science and Engineering, Shaanxi Normal University, Xian 710119, China †

ABSTRACT:

Site-selective growth of crystalline semiconductors on gold nanocrystals remains a great challenge because of the difficult control of both nucleation and growth dynamics as well as the easy agglomeration and deformation of gold nanocrystals at high temperatures of 4001000 C. Here we report a facile wet-chemistry route for the selective growth of crystalline ceria at the ends of gold nanorods (Au NRs) in the presence of a small amount of bifunctional K2PtCl4. Due to the smaller steric hindrance at the ends than at the side surface, K2PtCl4 may preferentially adsorb at the ends of Au NRs, triggering the autoredox reaction with the ceria precursor to obtain crystalline CeO2 at the ends. Notably, the surface of grown ceria is rich in oxygen vacancies (OVs) that facilitate the adsorption and activation of N2 molecules. The unique structure, the plasmon-induced hot carriers and the OVs make the obtained Au/end-CeO2 an excellent catalyst for nitrogen photofixation under near-infrared (NIR) illumination.

Triggering chemical reactions with solar energy offers exciting opportunities for the achievement of various challenging reactions under mild and environmentally friendly conditions. The major obstacle for the conversion of solar energy into chemical energy is the poor visible and NIR photocatalytic activity of traditional oxide semiconductor photocatalysts. An appealing route to solve these issues is the integration of oxide semiconductors with plasmonic metal nanocrystals, known as plasmonic photocatalysts.1,2 Plasmonic metal nanostructures (e.g., Au nanocrystals) can efficiently extend the photocatalytic activity of semiconductors to the visible and NIR regions through the hotelectron injection mechanism.3,4 Once hot electrons are injected into the semiconductor, holes left on Au nanocrystals should be consumed simultaneously, because the accumulation of holes will prevent the further injection and consequently lower the photocatalytic activity. The ideal plasmonic photocatalysts should have spatially-separated structures that allow holes and electrons to participate in the reaction simultaneously.510 For example, isotropic core@shell nanostructures bury the Au nanocrystal completely in the oxide shell, which makes the hot holes hardly accessible by reactant molecules.10 In contrast, the spatiallyseparated architecture allows hot holes and electrons to be freely

accessed by reactant molecules, resulting in a higher photocatalytic activity.5 Therefore, spatially-separated nanostructures of plasmonic photocatalysts are highly desired. Since the synthesis of spatially-separated nanostructures requires the site-selective nucleation and the growth of oxide semiconductors on plasmonic nanocrystals, it has remained challenging.1113 Janus Au/TiO2 nanostructures11,12 and (Au NR)/TiO2 nanodumbbell structure13 can be achieved by the controlled hydrolysis of titania precursors, but the crystalline phases of oxides are all amorphous which may jeopardize the photocatalytic activity due to the poor electron transport effiency.14 Up to date, selective coating of crystalline oxide semiconductors on the ends of Au NRs has not been reported, because the decomposition of oxide precursors generally requires high temperatures of more than 400 C,15 which not only makes the reaction dynamics difficult to control, but also causes Au NRs to deform and agglomerate. CeO2 as an n-type semiconductor is a good candidate for plasmonic photocatalysis, because it can form an appropriate Schottky barrier with Au nanocrystals to facilitate the hot-electron injection.3,16 Moreover, the ceria-based materials possess excellent catalytic capability due to the extensive presence of OVs on their surfaces.17 Benefiting from the tunable longitudinal plasmon wavelength,18 the large extinction cross-sections,18 and the preferential hot-electron generation on the ends,19 Au NRs are of particular interest as the core for the selective growth. In this study, we present a simple route for the site-selective growth of crystalline ceria on Au NRs at a moderate temperature. The K2PtCl4 plays a bifunctional role in the synthesis including (1) the selective growth through the preferential adsorption at two ends of Au NRs and (2) the activation of an autoredox reaction with the ceria precursor at the ends. The obtained hetero-nanostructures are Au NRs capped with crystalline ceria at the ends (i.e., Au/endCeO2). Such a spatial distribution makes hot electrons participate in the reduction reaction at the ends and hot holes be consumed at the exposed side surface, greatly enhancing the utilization efficiency of photocarriers. The unique structure, the plasmon effect and the OVs on the surface of ceria endow the Au/endCeO2 nanostructures with a high catalytic activity towards N2 photofixation under NIR light illumination.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1 illustrates the process for the selective coating of ceria onto a gold nanorod to produce a Au/end-CeO2 nanostructure. The pre-grown Au NR is stabilized with a cetyltrimethylammonium bromide (CTAB) bilayer. Because of the crystal structure and the curvature difference (see the mechanism of the selective nucleation at the ends of Au NRs in the Supporting Information), the molecular chains of CTAB at the ends are less dense in space than those on the side surface and thereby have smaller steric hindrance for other species to reach the Au NR.13,20 As a result, PtCl42- ions preferentially adsorb at two ends of Au NR. The ceria precursor (i.e., cerium acetate, Ce(AC)3) can be rapidly hydrolyzed into Ce(OH)3 when the temperature is higher than 60 C,21 and subsequently be oxidized by the pre-adsorbed PtCl42- according to the autoredox reaction to produce the nuclei for the further growth of CeO2. Consequently, the preferential adsorption of PtCl42- and the autoredox reaction between PtCl42- and Ce(OH)3 result in a site-selective nucleation and growth of CeO2, generating a Au/end-CeO2 heteronanostructure.

Figure 1. Schematic illustration of the synthesis process of a Au/end-CeO2 nanostructure. Three Au NR samples with different aspect ratios were used for the ceria growth (Figure S1). The representative transmission electron microscopy (TEM) images of the obtained Au/end-CeO2 nanostructures are shown in Figure 2a-c. The Au/end-CeO2 nanostructures exhibit uniform morphologies and narrow size distributions. The CeO2 is selectively grown at two ends of the Au NRs, leaving the side surface exposed. The thicknesses of the CeO2 layer in three samples are 12.9  1.8 nm, 15.0  1.8 nm and 12.5  1.6 nm, respectively. In addition, the TEM image taken in large areas (Figure S2) demonstrates a high yield (96%) of the Au/end-CeO2 nanostructures. Owing to the larger refractive index of CeO2 relative to water, the longitudinal plasmon resonance peaks of the Au NRs exhibit a redshift after the end-coating of CeO2 (Figure S1d).15,16

Figure 2. (ac) TEM images of the Au/end-CeO2 nanostructures produced from three different sized Au NRs. (d) HAADF-STEM

Page 2 of 5

image and the corresponding elemental maps. (e) HRTEM images of a single Au/end-CeO2 nanostructure. To examine the structure and chemical composition of the Au/end-CeO2 nanostructures, high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) image, elemental mapping, and high-resolution transmission electron microscopy (HRTEM) image were obtained (Figure 2de). The HAADF-STEM image and elemental mapping clearly reveal that each nanostructure contains an Au NR at the centre, with the elements Ce and O locating at two ends (Figure 2d), consistent with the elemental profiles (Figure S3). The element Pt can hardly be distinguished by elemental mapping due to their small amount. The presence of Pt is verified by energy-dispersive X-ray analysis (Figure S4) and HRTEM image (Figure 2e). The HRTEM displays that the CeO2 shell is composed of many small crystalline CeO2 nanosheets. The lattice fringes of three components validate their crystalline properties, consistent with the X-ray diffraction result (Figure S5).

Figure 3. High-resolution Ce 3d XPS spectrum (a) and EPR spectrum (b) of the Au/end-CeO2 nanostructures. To examine the element valence states and OVs in the Au/endCeO2, X-ray photoelectron spectroscopy (XPS) (Figure S6) and low-temperature electron paramagnetic resonance (EPR) analysis were performed. The high-resolution Ce 3d XPS spectrum exhibits eight strong peaks (Figure 3a), six of which can be assigned to three main and satellite peaks of Ce (IV) state, while the other two belong to the main peaks of Ce (III) state.16,22 The ratio of Ce (IV) to Ce (III) is calculated to be 85:15 by integrating the peak areas. This result suggests that the dominant form of Ce is Ce (IV). In addition, the presence of an EPR peak with a g value of 2.0042 indicates the formation of OVs caused by the Ce (III) state (Figure 3b).23 Thus, both XPS and EPR results confirm that the CeO2 component has abundant OVs on the surface. To gain a deep insight into the selective growth process, a series of control experiments were conducted (Figure S7S9). The K2PtCl4 amount is closely associated with the morphology of the obtained nanostructures (Figure S7). The morphology evolves from the self-nucleate CeO2 to the Au/end-CeO2 nanostructure, and then to the core@shell nanostructure when the K2PtCl4 amount increases from 0 to 200 L and further to 2000 L. In addition, the CTAB concentration (Figure S8) and proper precursor (Figure S9) are crucial to the selective growth behaviour. N2 fixation is industrially accomplished at high temperatures and high pressures because a large activation barrier (941 kJ/mol) should be overcome.24,25 Recent studies demonstrated that N2 molecules can be chemisorbed and activated by OVs and subsequently reduced to NH3 by the injected plasmonic hot electrons.2628 The spatially-separated structure and the presence of OVs make the Au/end-CeO2 nanostructures a preferable candidate for plasmon-induced NIR N2 photofixation. To examine the N2 photofixation performance, an Au NR and a fully coated Au@CeO2 core@shell nanostructure (Figure S8a) were used for comparison. Their longitudinal plasmon wavelengths were finely adjusted as closely as possible to the laser wavelength (Figure 4a). Plasmon-induced hot carrier generation was first examined through the detection of singlet oxygen under the laser

ACS Paragon Plus Environment

Page 3 of 5 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society illumination (Figure S10).15 The Au/end-CeO2 exhibits an excellent performance with a rate constant of 0.68  10-2 min-1, which is 8.9- and 5.0-fold compared with that of the bare Au NR and the core@shell nanostructures, suggesting that the spatiallyseparated structure facilitates the electronhole separation.

Figure 4. Plasmonic N2 photofixation under NIR illumination. (a) Extinction spectra of three photocatalyst nanostructures. (b) N2 photofixation rates of three photocatalysts. (c) Mechanism of N2 photofixation for the Au/end-CeO2 nanostructures. Ef, Fermi level; CB, conduction band; VB, valence band; OV, oxygen vacancy states. (d) Comparison of the hot carrier separation behaviors of the Au/end-CeO2 nanostructure with that of the core@shell nanostructure. The N2 photofixation activities were quantified by the produced amount of ammonia measured by the indophenol-blue method.27,29,30 The linear relationship between the absorption and the NH4+ concentration was pre-calibrated (Figure S11). The Au NR shows no NIR photocatalytic activity (Figure 4b, S12a). The concentration of NH3 increases linearly with the reaction time for the core@shell and the end-coated samples as the catalysts (Figure S12a). The N2 fixation rate of the Au/end-CeO2 nanostructure is 114.3 molh-1g-1 under the 808-nm laser illumination, which is 6.2-fold compared with that of the core@shell nanostructure. Control experiments reveal that NH3 results from the reduction of N2 instead of the contamination and the impurity (Figure S12b). A temperature of 56 C due to the photothermal effect is observed. However, the dominant contribution to the photocatalytic performance is believed to root in plasmonic hot carriers by unlocking reaction pathways and reducing the activation barriers.2 Stability tests indicate the excellent photostability of these nanostructures for NIR N2 photofixation (Figure S13). As a proof-of-concept, the photocatalytic performance of the Au/end-CeO2 nanostructures under 1 sun illumination was conducted and a N2 fixation rate of 25.6 molh-1g-1 was achieved (Figure S14). Based on the above results, we proposed a possible mechanism for the plasmon-induced NIR N2 photofixation. As shown in Figure 4c, the inherent OV states (i.e., Ce3+) on the surface of CeO2 act as the catalytic active sites that chemisorb and activate N2 molecules. Under the plasmon excitation, Au NRs absorb NIR light to generate hot electrons and hot holes. The produced hot electrons are injected into the CB of CeO2 across the Schottky barrier, followed by the transport to the OV states where the adsorbed and activated N2 is reduced into NH3. Meanwhile, the hot holes are consumed by the hole scavengers at the side of the Au NRs to complete the photocatalytic cycle. The spatial separation design of the Au/end-CeO2 nanostructures offers reaction sites for both reduction and oxidation, with the reduction

of N2 occurring on the surface of CeO2 and the consumption of hot holes on the side surface of the Au NRs (Figure 4d). In contrast, in the core@shell nanostructures, the Au NRs are buried inside, making hot holes be hardly accessed by CH3OH that results in the electronhole recombination (Figure 4d). Notably, a small activity obtained on the core@shell nanostructures may result from the presence of tiny gaps among the CeO2 nanosheets. In summary, we have designed and synthesized a new type of spatially-separated nanostructures by selectively coating crystalline ceria on the Au NRs using a wet-chemistry strategy at a moderate temperature. These nanostructures can be achieved with the assistance of bifunctional K2PtCl4 that can selectively adsorb on two ends of Au NRs to trigger the autoredox reaction of the ceria precursor. In addition, the presence of OVs in combination with the spatial separation structure makes the Au/end-CeO2 an excellent candidate for plasmon-induced NIR N2 photofixation. Further improvement of the NIR photocatalytic activity may be accomplished by either extending the light absorption with different Au/end-CeO2 mixtures or engineering with the co-catalysts. We believe that the proposed selective growth strategy will motivate the rational design of various spatial-separated plasmonic nanostructures for novel photocatalytic applications.

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website. Experimental details, additional results of structure characterizations.

AUTHOR INFORMATION Corresponding Author *[email protected]. *[email protected]. *[email protected]. ORCID Henglei Jia: 0000-0001-5882-1627 Ruibin Jiang: 0000-0001-6977-3421 Jianfang Wang: 0000-0002-2467-8751 Chun-yang Zhang: 0000-0002-8010-1981

Notes The authors declare no competing financial interest.

ACKNOWLEDGMENT This work was supported by the National Natural Science Foundation of China (Grant Nos. 21527811, 21735003, and 61775129), NSFC/RGC Joint Research Scheme (Ref. No. N_CUHK440/14), the Award for Team Leader Program of Taishan Scholars of Shandong Province, China, and Fundamental Research Funds for the Central Universities (GK201902001).

REFERENCES (1) Linic, S.; Christopher, P.; Ingram, D. B. Plasmonic-metal nanostructures for efficient conversion of solar to chemical energy. Nat. Mater. 2011, 10, 911921. (2) Zhou, L. N.; Swearer, D. F.; Zhang, C.; Robatjazi, H.; Zhao, H. Q.; Henderson, L.; Dong, L. L.; Christopher, P.; Carter, E. A.; Nordlander, P.; Halas, N. J. Quantifying hot carrier and thermal contributions in plasmonic photocatalysis. Science 2018, 362, 6972. (3) Clavero, C. Plasmon-induced hot-electron generation at nanoparticle/metal-oxide interfaces for photovoltaic and photocatalytic devices. Nat. Photonics 2014, 8, 95103.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(4) Jiang, R. B.; Li, B. X.; Fang, C. H.; Wang, J. F. Metal/semiconductor hybrid nanostructures for plasmon-enhanced applications. Adv. Mater. 2014, 26, 52745309. (5) Mubeen, S.; Lee, J.; Singh, N.; Krämer, S.; Stucky, G. D.; Moskovits, M. An autonomous photosynthetic device in which all charge carriers derive from surface plasmons. Nat. Nanotechnol. 2013, 8, 247251. (6) Weng, L.; Zhang, H.; Govorov, A. O.; Ouyang M. Hierarchical synthesis of non-centrosymmetric hybrid nanostructures and enabled plasmon-driven photocatalysis. Nat. Commun. 2014, 5, 4792. (7) Zheng, D. D.; Cao, X.-N.; Wang, X. C. Precise formation of a hollow carbon nitride structure with a Janus surface to promote water splitting by photoredox catalysis. Angew. Chem., Int. Ed. 2016, 55, 1151211516. (8) Naya, S.; Kume, T.; Akashi, R.; Fujishima, M.; Tada, H. Red-lightdriven water splitting by Au(core)CdS(shell) half-cut nanoegg with heteroepitaxial junction. J. Am. Chem. Soc. 2018, 140, 12511254. (9) Zhao, Q.; Ji, M. W.; Qian, H. M.; Dai, B. S.; Weng, L.; Gui, J.; Zhang, J. T.; Ouyang, M.; Zhu, H. S. Controlling structural symmetry of a hybrid nanostructure and its effect on efficient photocatalytic hydrogen evolution. Adv. Mater. 2014, 26, 13871392. (10) Jia, H. L.; Zhu, X.-M.; Jiang, R. B.; Wang, J. F. Aerosol-sprayed gold/ceria photocatalyst with superior plasmonic hot electron-enabled visible-light activity. ACS Appl. Mater. Interfaces 2017, 9, 25602571. (11) Seh, Z. W.; Liu, S. H.; Low, M.; Zhang, S.-Y.; Liu, Z. L.; Mlayah, A.; Han, M.-Y. Janus Au-TiO2 photocatalysts with strong localization of plasmonic near-fields for efficient visible-light hydrogen generation. Adv. Mater. 2012, 24, 23102314. (12) Seh, Z. W.; Liu, S. H.; Zhang, S.-Y.; Bharathi, M. S.; Ramanarayan, H.; Low, M.; Shah, K. W.; Zhang, Y.-W.; Han, M.-Y. Anisotropic growth of titania onto various gold nanostructures: synthesis, theoretical understanding, and optimization for catalysis. Angew. Chem., Int. Ed. 2011, 50, 1014010143. (13) Wu, B. H.; Liu, D. Y.; Mubeen, S.; Chuong, T. T.; Moskovits, M.; Stucky, G. D. Anisotropic growth of TiO2 onto gold nanorods for plasmon-enhanced hydrogen production from water reduction. J. Am. Chem. Soc. 2016, 138, 11141117. (14) Liu, G.; Yang, H. G.; Pan, J.; Yang, Y. Q.; Lu, G. Q.; Cheng, H.M. Titanium dioxide crystals with tailored facets. Chem. Rev. 2014, 114, 95599612. (15) Fang, C. H.; Jia, H. L.; Chang, S.; Ruan, Q. F.; Wang, P.; Chen, T.; Wang, J. F. (Gold core)/(titania shell) nanostructures for plasmonenhanced photon harvesting and generation of reactive oxygen species. Energy Environ. Sci. 2014, 7, 34313438. (16) Li, B. X.; Gu, T.; Ming, T.; Wang, J. X.; Wang, P.; Wang, J. F.; Yu, J. C. (Gold core)@(ceria shell) nanostructures for plasmon-enhanced catalytic reactions under visible light. ACS Nano 2014, 8, 81528162. (17) Paier, J.; Penschke, C.; Sauer, J. Oxygen defects and surface chemistry of ceria: quantum chemical studies compared to experiment. Chem. Rev. 2013, 113, 39493985. (18) Chen, H. J.; Shao, L.; Li, Q.; Wang, J. F. Gold nanorods and their plasmonic properties. Chem. Soc. Rev. 2013, 42, 26792724. (19) Brongersma, M. L.; Halas, N. J.; Nordlander, P. Plasmon-induced hot carrier science and technology. Nat. Nanotechnol. 2015, 10, 2534. (20) Wang, F.; Cheng, S.; Bao, Z. H.; Wang, J. F. Anisotropic overgrowth of metal heterostructures induced by a site-selective silica coating. Angew. Chem., Int. Ed. 2013, 52, 1034410348. (21) Song, S. Y.; Li, K.; Pan, J.; Wang, F.; Li, J. Q.; Feng, J.; Yao, S.; Ge, X.; Wang, X.; Zhang, H. J. Achieving the trade-off between selectivity and activity in semihydrogenation of alkynes by fabrication of (asymmetrical Pd@Ag core)@(CeO2 shell) nanocatalysts via autoredox reaction. Adv. Mater. 2017, 29, 1605332. (22) Zhang, F.; Wang, P.; Koberstein, J.; Khalid, S.; Chan, S.-W. Cerium oxidation state in ceria nanoparticles studied with X-ray photoelectron spectroscopy and absorption near edge spectroscopy. Surf. Sci. 2004, 563, 7482. (23) Nagarale, R. K.; Hoss, U.; Heller, A. Mixed-valence metal oxide nanoparticles as electrochemical half-cells: substituting the Ag/AgCl of reference electrodes by CeO2-x nanoparticles. J. Am. Chem. Soc. 2012, 134, 2078320787. (24) Jia, H.-P.; Quadrelli, E. A. Mechanistic aspects of dinitrogen cleavage and hydrogenation to produce ammonia in catalysis and organometallic chemistry: relevance of metal hydride bonds and dihydrogen. Chem. Soc. Rev. 2014, 43, 547564.

(25) Yang, J. H.; Guo, Y. Z.; Lu, W. Z.; Jiang, R. B.; Wang, J. F. Emerging applications of plasmons in driving CO2 reduction and N2 fixation. Adv. Mater. 2018, 30, 1802227. (26) Li, J.; Li, H.; Zhan, G. M.; Zhang, L. Z. Solar water splitting andnitrogen fixation with layered bismuth oxyhalides. Acc. Chem. Res. 2017, 50, 112121. (27) Yang, J. H.; Guo, Y. Z.; Jiang, R. B.; Qin, F.; Zhang, H.; Lu, W. Z.; Wang, J. F.; Yu, J. C. High-efficiency working-in-tandem nitrogen photofixation achieved by assembling plasmonic gold nanocrystals on ultrathin titania nanosheets. J. Am. Chem. Soc. 2018, 140, 84978508. (28) Hirakawa, H.; Hashimoto, M.; Shiraishi, Y.; Hirai, T. Photocatalytic conversion of nitrogen to ammonia with water on surface oxygen vacancies of titanium dioxide. J. Am. Chem. Soc. 2017, 139, 1092910936. (29) Greenlee, L. F.; Renner, J. N.; Foster, S. L. The use of controls for consistent and accurate measurements of electrocatalytic ammonia synthesis from dinitrogen. ACS Catal. 2018, 8, 78207827. (30) Zhu, D.; Zhang, L. H.; Ruther. R. E.; Hamers, R. J. Photoilluminated diamond as a solid-state source of solvated electrons in water for nitrogen reduction. Nat. Mater. 2013, 12, 836841.

ACS Paragon Plus Environment

Page 4 of 5

Page 5 of 5 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Table of Contents Graphic

ACS Paragon Plus Environment