Abstract
There is a pressing global need to increase the use of renewable energy sources and limit greenhouse gas emissions. Towards this goal, highly efficient and molecularly selective chemical processes that operate under mild conditions are critical. Plasmonic photocatalysis uses optically-resonant metallic nanoparticles and their resulting plasmonic, electronic, and phononic light-matter interactions to drive chemical reactions. The promise of simultaneous high-efficiency and product-selective reactions with plasmon photocatalysis provides a compelling opportunity to rethink how chemistry is achieved. Plasmonic nanoparticles serve as nanoscale ‘antennas’ that enable strong light–matter interactions, surpassing the light-harvesting capabilities one would expect purely from their size. Complex composite structures, combining engineered light harvesters with more chemically active components, are a focal point of current research endeavors. In this review, we provide an overview of recent advances in plasmonic catalysis. We start with a discussion of the relevant mechanisms in photochemical transformations and explain hot-carrier generation and distributions from several ubiquitous plasmonic antennae. Then we highlight three important types of catalytic processes for sustainable chemistry: ammonia synthesis, hydrogen production and CO2 reduction. To help elucidate the reaction mechanism, both state-of-art electromagnetic calculations and quantum mechanistic calculations are discussed. This review provides insights to better understand the mechanism of plasmonic photocatalysis with a variety of metallic and composite nanostructures toward designing and controlling improved platforms for green chemistry in the future.
1 Introduction
What common thread unites industries spanning plastics, pharmaceuticals, cosmetics, agricultural, and fertilizers? The answer is heterogeneous catalysis. Approximately 40 % of the world’s gross domestic product relies on heterogeneous catalysts to facilitate the synthesis of valuable chemical products on a massive scale [1]. Traditional thermal-driven catalysis utilizes active transition metals, such as platinum group metals, supported on metal oxides (e.g. Al2O3, MgO, CeO2, etc.). Reactions typically operate under high temperatures and pressures to satisfy the thermodynamic and kinetic constraints necessary for molecular conversion. The extreme conditions required for many reactions come with significant energy consumption, greenhouse gas emission, and additional requirements for separating reaction byproducts and recycling and regenerating the catalysts. Achieving precision chemistry, where reactions are simultaneously high-yield, product selective, and free from greenhouse-gas emissions, will be critical for a sustainable future.
Light-driven reactions, powered directly by sunlight or solar-driven LEDs, hold promise for such precision chemistry. Decades of progress in the field of optics have enabled the manipulation of light at the nanoscale, resulting in efficient light-harvesting materials which have recently been shown to enable new chemical pathways. The collective oscillation of free electrons at resonant conditions with nanoscale metal antennas, known as localized surface plasmon resonances (LSPR), offers the potential to reinvent photochemical transformations by accessing potential energy surfaces through non-equilibrium excited states [2]. Using these excited-state reaction pathways which are inaccessible in thermally-driven catalysis, it is possible to achieve higher reaction rates as well as enhanced chemical selectivity, as demonstrated in propylene epoxidation [3], selective acetylene hydrogenation [4], carbon dioxide reduction [5–9] and methanol steam reforming [10, 11]. Further, such chemical transformations can be triggered at far milder temperatures and pressures than traditional thermal-driven processes.
In recent years, there has been significant growth in the field of plasmonic chemistry, with numerous excellent review articles exploring the topic [12–16]. These articles cover various aspects of plasmonic chemistry, including computational catalyst design, novel materials and hybrid composites, and applications of potential industrially relevant reactions. In contrast to these contributions, we review recent developments in sustainable chemistry applications of some of the most greenhouse-gas emitting chemical industries (ammonia synthesis, hydrogen production) and efforts towards CO2 utilization. The chemical industry contributes more than 3.6 × 1013 kg of annual global greenhouse gas emissions, corresponding to 10 % of global emissions; we therefore believe the three reactions we cover are the most suitable to tackle a transition to a sustainable, circular economy [17, 18]. We systematically summarize the efficiency and mechanism of each catalyst, while providing a fundamental understanding from the angle of light–matter interaction and surface chemistry. Our review therefore bridges the gap between fundamental plasmonic chemistry and its next-step applications in sustainable heterogeneous catalysis.
In this review, we will describe recent progress in sustainable plasmonic chemistry, focusing on three key reactions – ammonia synthesis, hydrogen production and CO2 reduction – and how they can address issues such as energy storage, greenhouse gas emissions alleviation, and the development of a hydrogen economy. We start by briefly introducing the decay pathways and associated lifetimes of LSPRs to understand how they convert light energy into chemical energy. We describe the design of nanostructures for both strong light–matter interactions and energy transfer near the catalyst active centers. We briefly highlight the many excellent demonstrations of plasmon catalysis utilizing metals such as Pt, Ir, Pd, and then describe recent progress in utilizing sustainable materials to replace noble metals and rare materials. We also discuss recent progress in electromagnetic simulations and quantum mechanistic calculations. Finally, we will provide a brief outlook on the field and the questions which, from our perspective, must be addressed in the future.
2 General mechanism for photochemical transformation
Recent focus in plasmonic photocatalysis has been dedicated to unraveling the mechanisms driving chemical reactions in the presence of photoexcited plasmonic nanoparticles [2, 19]. The current understanding in the field is that resonant light–matter interactions, based on nanoparticle geometry and material, lead to the generation of excited charge carriers [20, 21]. These carriers possess a high level of chemical reactivity. Resonant excitation of the LSPR in a plasmonic nanoparticle causes near-field electromagnetic enhancements of several orders of magnitude, essentially ‘focusing’ light to sub-wavelength spaces (Figure 1A, i). Plasmons can subsequently dephase non-radiatively through surface-assisted Landau damping, an intrinsically quantum mechanical phenomenon which generates hot electron-hole pairs within 100 fs (Figure 1A, ii). Hot carriers can undergo further relaxation through electron-electron interactions, producing a quasi-Fermi-Dirac distribution within 1 ps (Figure 1A, iii). Photochemical transformation can be facilitated by hot carriers generated from the plasmon dephasing as well as those generated by direct interband transitions, and the dynamics and time scales of these processes are distinct. If the hot carriers are not extracted to surface adsorbates (or in the form of photocurrents for other applications), the kinetic energy of the hot carriers will be dissipated as heat to the surrounding medium (Figure 1A, iv).
![Figure 1:
Lifetime of plasmon due to electron-hole dephasing and associated hot carrier distributions. (A) The general process of plasmon decay: (i) excitation of the localized surface plasmon resonance (LSPR) results in an enhanced interaction cross-section with the incident light from the enhanced optical near-field. (ii) Non-radiative decay through Landau damping and pumping of non-equilibrium hot carriers. (iii) Hot-carrier thermalization through electron-electron scattering. The numbers of energetic carriers (also referred to as thermalized carriers) increases. (iv) Photothermal heating from electron-phonon interaction. (B) The energy distribution of the hot carriers (z axis) as the function of plasmon frequency (y axis) and hot carrier energy (x axis) in Al, Ag, Au, and Cu. The red part represents the hot carrier generation through phonon-assisted Landau damping (or intraband transition), and the blue part represents the direct photoexcitation of interband transition. Reproduced from ref. [22]. Copyright 2016 American Chemical Society.](/document/doi/10.1515/nanoph-2023-0149/asset/graphic/j_nanoph-2023-0149_fig_001.jpg)
Lifetime of plasmon due to electron-hole dephasing and associated hot carrier distributions. (A) The general process of plasmon decay: (i) excitation of the localized surface plasmon resonance (LSPR) results in an enhanced interaction cross-section with the incident light from the enhanced optical near-field. (ii) Non-radiative decay through Landau damping and pumping of non-equilibrium hot carriers. (iii) Hot-carrier thermalization through electron-electron scattering. The numbers of energetic carriers (also referred to as thermalized carriers) increases. (iv) Photothermal heating from electron-phonon interaction. (B) The energy distribution of the hot carriers (z axis) as the function of plasmon frequency (y axis) and hot carrier energy (x axis) in Al, Ag, Au, and Cu. The red part represents the hot carrier generation through phonon-assisted Landau damping (or intraband transition), and the blue part represents the direct photoexcitation of interband transition. Reproduced from ref. [22]. Copyright 2016 American Chemical Society.
For photochemical transformations, hot-carrier mediated processes are considered the most efficient pathway among all energy-transfer mechanisms [23, 24]. They additionally provide higher reactivity and chemical selectivity under mild conditions than traditional thermal-driven process [3, 25]. The most well-studied plasmonic metals include Au, Ag, Al, and Cu [13, 26–28]. Energy distributions of hot carriers calculated by first-principles calculations as a function of carrier energy and plasmon frequency are outlined in Figure 1B [22]. The indirect (red) and direct (blue) components represent phonon-assisted intraband transitions and interband transitions, respectively. Hot-electron-induced reactions are well studied using small Au and Ag nanoparticles (generally <20 nm in diameter) through their hot electrons generated after their dephasing due to Landau damping [7, 29], [30], [31], [32]. The relevant energy distribution can also be achieved through the interband transition of Al [33, 34]. Recent studies of hot holes in Au and Cu show that they can be extracted to drive oxidation reactions [35–39]. Other types of plasmonic materials can be used, such as doped semiconductors, perovskites, 2D materials, conductive polymers, etc. [40] Their properties and the design of nanostructures using these materials differ in some ways from metallic nanoparticles, and we refer the interested readers to these references for more information [41–44].
A plethora of opportunities exist in tuning plasmonic structures to optimize both light harvesting properties and chemical reactivity. The electromagnetic properties of the metallic nanoparticles can be modified by changing the size, shape, composition, and dielectric medium [29]. These changes to the electromagnetic properties impact the resonance frequencies of the systems and therefore the hot carrier generation. In addition to utilizing the plasmonic metals alone as both the light-harvester and catalytic center, combining them with more traditional catalytic materials, such as transition metals or metal oxides, through nanostructured designs like antenna-reactor, core–shell, and homogeneous alloying provides additional parameter space to explore in pursuit of effective materials. In the case of alloying, a large focus has been placed on combating the inherent trade-off between plasmonic and chemically active materials. Based on ground-state electronic structure, it is often the case that noble metals (Au, Ag), which are high-performing light absorbers, have lower intrinsic reactivity than many transition group precious metals (Pd, Pt, Ir, etc.). Combining these materials can synergistically provide advantages from both groups. Meanwhile, recent work seeks to find alloys that can utilize more sustainable materials instead, such as Al, Fe, and Cu. In addition to the design of sustainable materials systems, the bulk of research in the field seeks to address chemical reactions that align with sustainability concerns, particularly in the areas of fertilizer production, greenhouse gas remediation, and high-volume chemical feedstocks.
3 Applications for sustainable chemistry
3.1 Ammonia synthesis (N2 reduction)
The Haber-Bosch Process is the traditional method of producing ammonia for fertilizer and emblematic of the challenge researchers must address to move the chemical industry into a sustainable future. Global population growth has been enabled specifically because of the invention of the ammonia synthesis process, contributing to ∼50 % of world population. Historically, around 78 % of the ammonia produced by the Haber-Bosch process is used in agriculture, as fertilizer to feed the world [45, 46]. Yet, despite its societal impact, the Haber-Bosch process is also responsible for a considerable carbon footprint. Traditionally, Fe-based catalysts supported on oxide are used under high temperature (300–500 °C) and pressure (100 atm) to convert N2 and H2 into NH3 [47, 48]. The process accounts for around 2 % of total energy from the chemical industry, and consumes 50 % of hydrogen production within the sector, leading to 3 % of annual global greenhouse gas emissions [49, 50]. There are few chemical reactions as important to the world which so critically necessitate innovation.
The fixation of dinitrogen species onto the surface of active centers is a crucial step in the reaction. The electronic structure of transition metals such as Os, Ru, and Fe makes them promising candidates to provide the active surface for this step. For instance, the first and second generations of the Haber-Bosch catalysts, which are based on Fe and Ru, respectively, rely on the Fe C7 and Ru B5 active sites for the activation of dinitrogen species [14, 51]. Plasmonic antennas offer a viable approach for improving the activation and fixation of dinitrogen species through the strong optical near-field enhancement, facilitating electronic transitions of the adsorbates by hot carriers, and traditional equilibrium and thermodynamic effects from photothermal conversion of light energy.
Early photocatalytic studies of ammonia synthesis show that light irradiation can activate N2 at milder conditions than the industrial Haber-Bosch process. Zeng, H., et al. demonstrated a proof of concept for photochemical ammonia synthesis by using Au-Os nanoparticles created via sputtering on a glass substrate. Gas-phase ammonia synthesis (N2 + 3H2 → 2NH3) is achieved at room temperature and pressure using 200 mW/cm2 irradiation from a solar simulator with a 450 nm cutoff filter. The study proposes that plasmons are influential to the process, since the extinction spectra of the sample matches with wavelength-dependent reactivity measurements (Figure 2A). This study represents one of the initial demonstrations of using a thin layer of Os nanoparticles as the active catalyst to drive the reaction photochemically. It is noteworthy that Os was the first transition metal historically used for traditional ammonia synthesis [52]. In an aqueous phase (N2 + 6H+ + 6e− → 2NH3), Hu, C., et al. showed that Au-Ru nanoparticles with multiple bimetallic ratios under Xenon lamp illumination at 400 mW/cm2, 2 bar pressure, and room temperature can facilitate the associative alternating hydrogenation of dinitrogen species when activating the surface plasmon resonance. The research team demonstrated the plasmon-driven reactivity of the system through the linear relationship between the reactivity and the input power (Figure 2B). Moreover, their first-principle model provided mechanistic insights into the necessity of a strong electric near-field for the activation of dinitrogen species onto the Ru center [53]. Au nanoparticles on various semiconductors, such as TiO2, CeO2, and g-C3N4, achieve enhanced visible light absorption as well as charge separation at the metal-semiconductor interface for nitrogen fixation. Semiconductor materials with appropriate band structure and suitable active surfaces are inherently effective catalysts for nitrogen fixation under UV illumination. Incorporating plasmon-mediated visible light absorption with active oxygen or nitrogen defects can augment ammonia fixation, providing an additional avenue for promoting dinitrogen fixation that is distinct from traditional transition metal-based catalysts [54–57]. Supporting Au nanoparticles in the metal-organic framework (MOF) UiO-66 increases permeability owing to the large surface area of the matrix, better enabling the Au nanoparticles to activate N2 on their active sites. The MOF support provided excellent stability for over 24 h, a significant improvement compared to only 8 h when the same Au was supported on other metal oxides (Figure 2C). Traditionally, the Au surface of the MOF is a poor active center for the activation of dinitrogen species. However, the unique Au-MOF interface, which exhibits an enhanced surface area, displayed extraordinary reactivity and a decrease in light-induced activation [58]. There are also several studies utilizing Ru catalysts which are similar to industrial catalysts in components and structure, but now the plasmonic photothermal effect is synergistically employed to drive gas-phase ammonia synthesis. Li, X., et al. reported the LED illumination of a conventional Ru-Cs/MgO catalyst and utilized the temperature gradient caused by the plasmonic photothermal effect (Figure 2D) to achieve a reaction rate of 4464 μmol g−1 h−1 at 4.7 W/cm2, detecting the gas-phase conversion of N2 to NH3 in a mass spectrometer [59]. Mao, C., et al. reported a K/Ru/TiO2−x H x catalyst at 360 °C under 300 W Xe lamp irradiation and reaching 112.6 μmol g−1 h−1(Figure 2E) [60]. The same research team also investigated a Fe-based nano-necklace/TiO2−x H x and achieved ammonia concentration of 1939 ppm at 1 atm and 19,620 ppm at 10 atm [61]. Hou, T., et al. developed a porous Cu96Fe4 nanostructure and achieved 342 μmol g−1 h−1 at 250 mW/cm2 with the full spectrum of their Xenon lamp, and 242 μmol g−1 h−1 with a >400 nm filter at 200 mW/cm2. The catalyst also displays good stability during 10 cycles (Figure 2F) [62]. Table 1 summarizes the recent progress on N2 reduction. Taken together, these results show that plasmons can drive ammonia synthesis at milder temperatures and pressures than the Haber-Bosch process. Yet, further work is needed to regulate the dinitrogen activation towards the dissociative pathways and therefore boost its efficiency towards industrial scales. To identify avenues for controlling the associative hydrogenation of dinitrogen to distal and dissociation pathways, it is necessary to perform in-situ measurements of intermediate steps. Furthermore, high-resolution, and even atomic-resolution electron microscopy is essential to determine the active centers and their interaction evolution during the reaction. Theoretical studies are also crucial for gaining insights into the design of nanostructures, Particularly for understanding how excited-states can potentially lower the activation barrier of dinitrogen species.
![Figure 2:
Plasmonic nitrogen fixation. (A) The wavelength-dependent reactivity (blue open dots and dashed curve) is well-correlated to the extinction spectrum (pink solid line) on a Au-Os nanoparticles for gas-phase ammonia synthesis (N2 + 3H2 → 2NH3). Adapted from ref. [52] with permission. Copyright 2014 Elsevier B.V. (B) Ammonia synthesis in presence of water (N2 + 6H+ + 6e− → 2NH3) on a AuRu core-antenna nanostructure under full-spectrum irradiation at 2 atm N2. Adapted from ref. [53] with permission. Copyright 2019 American Chemical Society. (C) Stability test of ammonia synthesis on Au in presence of water on UiO-66 MOF. Adapted from ref. [63] with permission. Copyright 2019 American Chemical Society. (D) The plasmonic gas-phase ammonia synthesis with a commercial Ru-Cs/MgO catalyst driven by the thermal gradient. Adapted from ref. [59]. Copyright 2019 American Chemical Society. (E) Plasmonic photothermal gas-phase ammonia synthesis with a K/Ru/TiO2−x
H
x
catalyst. Adapted from ref. [60] with permission. Copyright 2017 Elsevier B.V. (F) Stability of a porous plasmonic CuFe for nitrogen fixation with 1 atm N2 in water. Adapted from ref. [62] with permission. Copyright 2020 American Chemical Society.](/document/doi/10.1515/nanoph-2023-0149/asset/graphic/j_nanoph-2023-0149_fig_002.jpg)
Plasmonic nitrogen fixation. (A) The wavelength-dependent reactivity (blue open dots and dashed curve) is well-correlated to the extinction spectrum (pink solid line) on a Au-Os nanoparticles for gas-phase ammonia synthesis (N2 + 3H2 → 2NH3). Adapted from ref. [52] with permission. Copyright 2014 Elsevier B.V. (B) Ammonia synthesis in presence of water (N2 + 6H+ + 6e− → 2NH3) on a AuRu core-antenna nanostructure under full-spectrum irradiation at 2 atm N2. Adapted from ref. [53] with permission. Copyright 2019 American Chemical Society. (C) Stability test of ammonia synthesis on Au in presence of water on UiO-66 MOF. Adapted from ref. [63] with permission. Copyright 2019 American Chemical Society. (D) The plasmonic gas-phase ammonia synthesis with a commercial Ru-Cs/MgO catalyst driven by the thermal gradient. Adapted from ref. [59]. Copyright 2019 American Chemical Society. (E) Plasmonic photothermal gas-phase ammonia synthesis with a K/Ru/TiO2−x H x catalyst. Adapted from ref. [60] with permission. Copyright 2017 Elsevier B.V. (F) Stability of a porous plasmonic CuFe for nitrogen fixation with 1 atm N2 in water. Adapted from ref. [62] with permission. Copyright 2020 American Chemical Society.
Summary of ammonia synthesis literature.
Nanostructure | Light | Power | Ammonia | Quantum | Mechanism | Reference |
---|---|---|---|---|---|---|
source | density | rate | efficiency | number | ||
Gas phase reaction: N2 + 3H2 → 2NH3 | ||||||
Au-Os/substrate | Solar simulated (>450 nm) | 200 mW/cm2 | 265 μmol gOs −1 h−1 | Au plasmon + Os active center | [52] | |
Ru-Cs/MgO | White and blue LED | 0–4.7 W/cm2 | 4500 μmol g−1 h−1 | Light-induced thermal gradients in ruthenium catalysts significantly enhance ammonia production | [59] | |
K/Ru/TiO2−x H x | 300 W Xenon lamp | 112.6 μmol g−1 h−1 | Enhanced hydrogen spillover and energy confinement at the interface | [60] | ||
Fe nano-necklace/TiO2−x H x | 5–10 W/cm2 | 1, 939 ppm (1 atm); 19,620 ppm (10 atm) | Dual temperature zone from the local heating of Fe | [61] | ||
N2 fixation in presence of water (N2 + 6H+ + 6e−→2NH3) | ||||||
Au-Ru core antenna | Xenon lamp at the full spectrum | 400 mW/cm2 | 101.4 μmol g−1 h−1 (2 atm N2) | 0.21 % at 350 nm | Near electric field on the surface of Ru | [53] |
Au/UiO-66 | Xenon lamp with 400 nm long pass filter | 100 mW/cm2 | 360 μmol g−1 h−1 (1 atm N2) | 0.59 % at 520 nm | Large surface area for gas adsorption | [58] |
Cu96Fe4 | Xenon lamp | 250 mW/cm2 (320–780 nm) | 342 μmol g−1 h−1; | 0.13 % at 535 nm | Valence state and coordination number of Fe increase | [62] |
200 mW/cm2 (400–780 nm) | 242 μmol g−1 h−1 | |||||
N2 fixation in presence of water and a sacrificial agent | ||||||
Au/TiO2 nanosheets | Xenon lamp at 420 nm cutoff filter | 78.6 μmol g−1 h−1 | 0.82 % at 550 nm | Oxygen vacancy for N2 chemisorption | [54] | |
Au nanorod/CeO2 | 808 nm laser | 8 W/cm2 | 114.3 μmol g−1 h−1 | Oxygen vacancy for N2 chemisorption | [55] | |
Broadband | 100 mW/cm2 | 25.6 μmol g−1 h−1 | ||||
Au embedded in hollow carbon nitride sphere | Xenon lamp at 420 nm cutoff filter | 783.4 μmol g−1 h−1 | Nitrogen vacancy for nitrogen activation | [57] | ||
Au/g-C3N4 | Xenon lamp at 420 nm cutoff filter | 93 μmol g−1 h−1 | Nitrogen vacancy for nitrogen activation | [56] |
3.2 H2 production
H2 is not only a ubiquitous reducing agent for a multitude of reactions in heterogeneous catalysis and organic/pharmaceutical synthesis, but also a promising carbon-free fuel for vehicles and a high-density storage technology [64]. Many renewable applications try to split water into hydrogen and oxygen via photocatalysis, electrocatalysis, or photoelectrocatalysis [65–68]. The first instance of hydrogen evolution through direct light-driven water splitting dates back to 1971, when Fujishima and Honda discovered that UV irradiation on the surface of TiO2 enabled the direct photocatalysis of H2O into H2 and O2 [69]. Since then, scientists have explored various strategies to engineer semiconductor materials to achieve high yields of green hydrogen from solar energy. To achieve solar-to-hydrogen efficiency that is competitive with other hydrogen generation methods, a theoretical energy efficiency of 10 % would require a single absorber with a band gap smaller than approximately 530–600 nm [70]. The first milestone was achieved by Moskovits and co-workers, who used a Au-TiO2-Pt/Co thin-film device for photoelectrochemical water splitting to demonstrate the proof of concept that plasmon-induced hot carriers can be extracted to drive water splitting [71, 72]. Recent advancements have focused mainly on the fundamental aspects of extracting hot electrons or hot holes to facilitate both quantum efficiency and energy efficiency [73–75]. Many demonstrations of utilizing composite heterostructure semiconductors with plasmonic antennas to enhance visible light absorption are discussed in detail in ref. [15, 76]. Challenges such as constructing appropriate multiple Schottky junctions between metal and semiconductor for efficient charge separation and understanding the multiple electron coupling process in aqueous-phase reactions need to be further investigated.
Several reactions are used to produce hydrogen in industry, including ammonia decomposition, cracking of fossil fuels, methane reforming, and the Claus process. These methods generally require harsh operating conditions (high temperatures and pressures with transition metal catalysts) [77]. Recent progress in plasmonic photocatalysis suggests that non-thermal hot-carrier-driven reactions can be used for the production of hydrogen from a variety of sources with improved selectivity and lower activation barriers at mild conditions under visible light irradiation.
Zhou, L., et al. reported Cu-Ru surface alloy antenna-reactors for ammonia decomposition (2NH3 → 2N2 + 3H2) [78]. Photocatalytic reactivity under white light illumination on the surface of Cu-Ru can be 1–2 orders of magnitudes higher than the reactivity from thermocatalysis at comparable temperatures. The energy efficiency for converting from light to chemical energy can be as high as 18 %. The apparent activation energies of the reaction under different wavelengths and power densities of light can decrease from 1.2 eV to 0.2 eV (Figure 3A), suggesting that hot carriers can help tailor the rate-determining step of this reaction, which shifts from the associative desorption of N2 in thermocatalyis to the N–H bond scission steps, therefore increasing the activity and reducing the activation barrier. Using the same synthetic techniques, the research team also compared Cu-Fe and Cu-Ru under similar light illumination conditions and found comparable performance. Originally Fe is famous for the strong adsorption with nitrogen species and can form FeN x , thereby deactivating the active centers. However, hot carriers can remove the surface N x species and facilitate the desorption of NH3, turning Fe into an active center for photocatalytically decomposing the NH3. Additionally, the authors were able to change the light source from a laser to LED while scaling the reaction by 3 orders of magnitude (Figure 3B) and maintaining an energy efficiency of more than 10 % [79].
![Figure 3:
Plasmonic gas-phase catalysis for H2 generation. (A) Apparent activation energies under different wavelengths and power densities on a Cu-Ru surface alloy antenna reactor for NH3 decomposition (2NH3 → 2N2 + 3H2). Adapted from ref. [78] with permission. Copyright 2018 The American Association for the advancement of science. (B) 3D reactor to scale up the Cu-Ru and Cu-Fe catalysts in gram scale for ammonia decomposition with LED light. Adapted from ref. [79] with permission. Copyright 2022 The American Association for the advancement of science. (C) Stability test of CuRu for methane dry reforming (CH4 + CO2 → 2CO + 2H2). Adapted from ref. [80] with permission. Copyright 2020 Springer Nature. (D) Production rate and selectivity evolution with varying temperature of Au-Pt/P25 for photothermal catalysis. Adapted from ref. [81] with permission. Copyright 2022 Elsevier Inc. (E) Reaction mechanism of Au on SiO2 for direct photocatalytic and thermocatalytic H2S decomposition. Adapted from ref. [31] with permission. Copyright 2022 American Chemical Society. (F) In-situ FTIR measurement to detect the intermediate species during photochemical transformation on CuZn alloy for methanol steam reforming (CH3OH + H2O → 3H2 + CO2). Adapted from ref. [10] with permission. Copyright 2021 American Chemical Society.](/document/doi/10.1515/nanoph-2023-0149/asset/graphic/j_nanoph-2023-0149_fig_003.jpg)
Plasmonic gas-phase catalysis for H2 generation. (A) Apparent activation energies under different wavelengths and power densities on a Cu-Ru surface alloy antenna reactor for NH3 decomposition (2NH3 → 2N2 + 3H2). Adapted from ref. [78] with permission. Copyright 2018 The American Association for the advancement of science. (B) 3D reactor to scale up the Cu-Ru and Cu-Fe catalysts in gram scale for ammonia decomposition with LED light. Adapted from ref. [79] with permission. Copyright 2022 The American Association for the advancement of science. (C) Stability test of CuRu for methane dry reforming (CH4 + CO2 → 2CO + 2H2). Adapted from ref. [80] with permission. Copyright 2020 Springer Nature. (D) Production rate and selectivity evolution with varying temperature of Au-Pt/P25 for photothermal catalysis. Adapted from ref. [81] with permission. Copyright 2022 Elsevier Inc. (E) Reaction mechanism of Au on SiO2 for direct photocatalytic and thermocatalytic H2S decomposition. Adapted from ref. [31] with permission. Copyright 2022 American Chemical Society. (F) In-situ FTIR measurement to detect the intermediate species during photochemical transformation on CuZn alloy for methanol steam reforming (CH3OH + H2O → 3H2 + CO2). Adapted from ref. [10] with permission. Copyright 2021 American Chemical Society.
Methane can be utilized to extract synthesis gas (carbon monoxide and hydrogen) and has gained increased attention in light of abundant methane sourced from shale gas and tight oil [82]. Methane enables hydrogen production from direct cracking, steam reforming by reacting with water, dry reforming with CO2, and oxidation. Dry reforming is the most environmentally-friendly strategy as it reforms two greenhouse emission gases into CO and H2 but it generally requires high operating temperatures (700–1000 °C). Zhou., L., et al. incorporated a single atom Ru into a Cu nanoparticle by the co-precipitation method and were able to drive methane dry reforming (CH4 + CO2 → 2CO + 2H2) under visible light irradiation with decent stability over 50 h without coking [80]. Several ratios of Cu-Ru were tested. However, most of them were deactivated due to coking on the surface, as shown in Figure 3C. Only Cu19.8Ru0.2 and Cu19.9Ru0.1 exhibited both reactivity and selectivity for over 20 h. The authors further confirmed that Ru existed in the form of a single-atom feature by utilizing infrared CO adsorption measurements combined with a DFT + D3 calculation. By comparing the reactivity and selectivity of thermocatalysis, they suggested that the plasmon-induced desorption assists in tailoring the pathway, and the single-atom Ru center feature prevents the adjacent carbon landing sites, thereby reducing coking. Furthermore, Zhang., Z., et al. reported a reaction rate of 85.38 mmol gcat −1 h−1 and 201.92 mmol gcat −1 h−1 for H2 and CO, respectively, using the plasmonic photothermal effect on their Pt-Au/P25 composite catalysts [81]. The results presented in this study demonstrate a 3-fold increase in reactivity compared to the reaction in the absence of light. Furthermore, the reactivity and selectivity were observed to be stable for over 15 h. The selectivity was higher at higher temperatures for both photo- and thermo-catalysis, indicating the dominance of the photothermal effect in the reaction (Figure 3D). Additionally, the authors used in-situ FTIR measurements to identify intermediate species of CH x , HCOO*, and H* on the surface of their catalysts. The HCOO* species was found to be the most important in driving the reaction towards the production of CO and H2 by decomposition. Based on these findings, a detailed mechanism was proposed in which plasmon-induced hot carriers on the surface of TiO2 help convert CH4 into CH x , while the Pt surface cleaves CO2 and spillovers the oxygen to the interface to facilitate the formation of HCOO*. Takami., D., et al. reported the activation of the reaction at the low temperature of 200 °C over a plasmonic Ni/Al2O3 catalyst, maintaining a high conversion of CH4 and CO2 (20 %) [83]. They recognize that the metallic Ni that support LSPR is the key to maintain the high conversion at low temperatures.
Other avenues for sustainable hydrogen production from plasmonic photocatalysis include hydrogen sulfide decomposition (H2S → H2 + S) and methanol steam reforming (CH3OH + H2O → 3H2 + CO2). Lou., M., et al. used plasmonic Au nanoparticles supported on silica without any external heating source to catalyze one-step direct H2S decomposition, a potential alternative for the industrial two-steps Claus process to obtain sulfur and H2 (Figure 3E) [31]. The results of ground and excited-state first-principle calculations, combined with partial-pressure dependent reactivity and micro-kinetic analysis, revealed that plasmon-induced hot electrons on the surface of Au nanoparticles shift the rate-determining steps from H–S* bond scission in dark conditions to S2 associative adsorption under light irradiation. Luo., S., et al. reported a H2 production rate of 328 mmol gcat −1 h−1 for methanol steam reforming on a plasmonic CuZn alloy, with a light energy conversion efficiency of 1.2 % at 7.9 suns irradiation from a solar simulator [10]. The in-situ diffuse reflective infrared Fourier transform spectra (DRIFTS) confirmed the key intermediate species on the copper-based catalysts are CH x O and HCOOH. The light illumination helps increase the concentration of the intermediates, thereby improving reactivity (Figure 3F). Gas-phase H2 production publications are summarized in Table 2. Incorporation of these state-of-art experiments into actual industrial production requires parameter optimization and large-scale photoreactor design. Indeed, several industries and start-ups are now addressing 3D design of reactors to better utilize all the photons for photochemical transformations and to improve flow of the gas to efficiently penetrate the reactor. Although the high-efficiency conversion of H2 generation developed so far in the field seems to facilitate the possibility of industrial-level production, fundamental studies on understanding the dynamics of hot carriers and how to extract them into chemistry are essential to increase quantum and energy efficiency. Moreover, renewable light sources such as solar energy have a rather complex wavelength distribution, making the study on extracting them more challenging than the exciting results achieved using monochromatic light sources such as lasers and LEDs.
Summary of gas-phase photocatalytic H2 production.
Nanostructure | Light source | Power density | H2 rate | Quantum/energy efficiency | Highlights of the work | Reference number |
---|---|---|---|---|---|---|
Ammonia decompostion: N2 + 3H2 → 2NH3 | ||||||
Cu-Ru surface alloy | Supercontinuum laser with a bandpass filter | 9.6 W/cm2 | 1200 μmol g−1 s−1 | 18 % (energy); 33.5 % (quantum) | Hot-electron tailored to the N–H bond scission. Activation energies range from 1.2 to 0.2 eV under illumination. | [78] |
Cu-Fe | Supercontinuum visible laser | 200 mW | 466 μmol g−1 s−1 | Photoreactor with LED light to scale up the reaction. | [79] | |
470 nm LED | 180 W | 14 g/day | 15.6 % (energy) | |||
Methane dry reforming: CH4 + CO2 → 2CO + 2H2 | ||||||
Cu-Ru | Supercontinuum laser | 19.2 W/cm2 | 550 μmol g−1 s−1 | 15 % (energy) | Single-atom Ru sites help alleviate the coking and long-term stability | [80] |
Au-Pt/P25 | Xenon lamp | 4.6 W/cm2 | 201.92 mmol g−1 h−1 | In-situ FTIR for the mechanism study | [81] | |
Ni/Al2O3 | Xenon lamp | 1.2 mmol h−1 | 20 % conversion rate under 473 K low temperature | [83] | ||
H2S decomposition: H2S → H2 + S | ||||||
Au/SiO2 | Supercontinuum visible laser | 13 W/cm2 | 5 μmol g−1 s−1 | 1.8 % (quantum) | One-step direct decomposition. Light-induced rate-determining step shift | [31] |
Methanol steam reforming: CH3OH + H2O → 3H2 + CO2 | ||||||
CuZn alloy | Solar simulator AM 1.5 G | 788 mW/cm2 | 328 mmol g−1 h−1 | 1.2 % (energy); 6.5 % (quantum) | Zn serves atoms serve as active sites and charge transfer center | [11] |
3.3 CO2 reduction
Reaction pathways with multiple electron transfer processes can potentially be used to reduce CO2 in the presence of water or H2 to turn CO2 into value-added chemical products [84]. The metal-molecule interface of the chemical transformation can be modified via optical near-fields, hot carrier generation and multiplication [85], localized photothermal effects and gradients [86, 87], and the construction of active centers. Control of these parameters allows for coupling optical energy into different products with high selectivity, allowing for tailor-made catalysts that target particular chemical reaction pathways.
Yu., S., et al. investigated the mechanism of C–C coupling on the surface of Au with CO2 in the presence of H2O. The selectivity was highly tailored by both the wavelength and power density of the incident continuous wave (CW) lasers: High-intensity resonant illumination of interband transitions at 488 nm showed preferred selectivity to ethane (corresponding to a multi-electron harvesting process) while excitation at 532 nm pushed the reaction towards methane production [7]. The team found that the observed selectivity trend can be attributed to the nature of higher light intensity and photon energy, which can result in a multi-carrier process. They also observed that the reactivity and selectivity follow the Poisson statistics of electron harvesting. The team also investigated the same reaction on Ag plasmonic nanoparticles at a single-particle level by employing in-situ Raman spectroscopy. They detected C1–C4 product evolutions with multiple intermediate species with visible light irradiation at mild conditions [6, 88]. Trace amounts of rich catalog of multiple carbon species were generated through a multi-photon excitation process during plasmon decay. However, the detailed mechanism of this process and how to regulate the selectivity towards long carbon-chain products require further investigation and development. Abundant reports have been published of hybrid plasmonic nanoparticles with semiconductor materials or MOFs to enhance the light–matter interaction and regulate the selectivity of the major products towards CH4 [89, 90], CO [91–93], CH3OH [94], HCOOH [63, 95, 96], and C2H6 [97], among others, by taking advantages of the unique electronic structure and vacancy sites on the metal oxide semiconductor or the large surface area of MOF. Understanding the mechanism and manipulating multi-electron processes through nanostructure engineering remains a future challenge of the field given the complex possible reaction pathways.
Gas-phase CO2 hydrogenation is another approach to use renewable light energy to produce valuable chemical products. Rather than the multiple electron transfer process as in the presence of water, gas-phase CO2 hydrogenation mainly has two reaction pathways under room pressure: CO2 methanation (CO2 + 4H2 → CH4 + 2H2O) and reverse water-gas shift reaction (CO2 + H2 → CO + H2O). Zhang., X., et al. observed the non-thermal effect on Rh/Al2O3 catalysts under visible LED illumination driving reaction towards CH4 with selectivities >86 % and >98 % under illuminations of blue (400 nm) and ultraviolet (365 nm) light, respectively [9]. Compared to thermocatalysis, plasmonic photocatalysis can selectively activate the C-O bond in CHO* intermediates by hot electrons with an apparent activation energy decrease (Figure 4A). The authors also reported a mechanistic study of a Rh/TiO2 catalyst, showing the selective methanation of CO2 through synergistic photothermal and hot-carrier effects (Figure 4B) [98]. The photothermal effect and the plasmon-mediated non-thermal effect (Figure 4B, iii) can be distinguished and quantified by measuring the top and bottom temperatures of the catalyst bed and combining them with a model. In addition, their DFT calculation found that CHO* is a critical intermediate for CO2 methanation in the presence of H2. However, increasing the hydrogen partial pressure would inhibit the injection of hot electrons into the orbital of CHO*. Robatjazi., H., et al. designed a sustainable aluminum-based core–shell structure of Al@MIL-53 and Al@Cu2O to drive the reverse water-gas shift reaction by increasing the physisorption of the gas, and extracting hot electrons from Al, respectively (Figure 4C and D) [5, 99]. Al@MIL-53 enables up to eight times larger CO2 uptake than Al nanocrystals, resulting in a four times higher reactivity towards the reverse water-gas shift reaction (shown in Figure 4C). In Al@Cu2O, the small Schottky barrier between Al and Cu2O (shown in Figure 4D) facilitates the separation of hot carriers generated from the Al core. This results in a regulation of selectivity towards the reverse water-gas shift reaction under visible-light illumination. Additionally, this boosts the EQE to two orders of magnitude higher than bare Cu2O and one order of magnitude higher than Al nanoparticles. Fu., G., et al. designed a Rh/Al photothermal catalyst from colloidal synthesis with Al powders and RhCl3 precursor, capable of regulating CO2 hydrogenation towards methanation with nearly 100 % selectivity under a concentrated simulated solar irradiation at 11.3 W/cm2, resulting in a 550 mmol g−1 h−1 reaction rate [8]. The in-operando Temperature-Programmed FTIR analysis identified the surface CO species on Rh is the key intermediate steps to drive the reaction towards the methanation process (Figure 4E). Table 3 outlines the relevant citations for gas-phase CO2 hydrogenation reactions. In our opinion, further research should focus on resolving the multiphoton and electron processes under high light intensity and photon energy in both aqueous and gas-phase CO2 reduction. Additionally, constructing active centers with different electronic structures for the various carbon-related species to accumulate and desorb can help regulate the selectivity of CO2 conversion into valuable molecules. Even for extensively studied gas-phase reactions such as reverse water-gas shift and methanation, understanding the time resolution of the intermediate steps is important to determine how to optimize the lifetime of plasmon excitation and decay to improve conversion.
![Figure 4:
Plasmonic gas-phase CO2 hydrogenation. (A) Activation energies and selectivity of CO2 hydrogenation towards methane on Rh/Al2O3 catalysts are tailored by light compared to the reaction in the dark. Adapted from ref. [9] with permission. Copyright 2017 Springer Nature. (B) CO2 methanation reaction on Rh/TiO2 catalysts utilizing both thermal and non-thermal effects with the temperature gradient in the catalyst bed. Adapted from ref. [98] with permission. Copyright 2018 American Chemical Society. (C) Al@MIL-53 for enlarged surface area for gas uptake and enhanced reverse water-gas shift reaction. Adapted from ref. [99] with permission. Copyright 2019 The American Association for the advancement of science. (D) Schematic of hot-electron induced chemistry at Al-Cu2O interface for reverse water-gas shift reaction. Adapted from ref. [5] with permission. Copyright 2017 Springer Nature. (E) In-operando FTIR measurement of reaction steps under different temperatures for photothermal CO2 methanation on a Rh/al catalyst. Adapted from ref. [8] with permission. Copyright 2021 American Chemical Society.](/document/doi/10.1515/nanoph-2023-0149/asset/graphic/j_nanoph-2023-0149_fig_004.jpg)
Plasmonic gas-phase CO2 hydrogenation. (A) Activation energies and selectivity of CO2 hydrogenation towards methane on Rh/Al2O3 catalysts are tailored by light compared to the reaction in the dark. Adapted from ref. [9] with permission. Copyright 2017 Springer Nature. (B) CO2 methanation reaction on Rh/TiO2 catalysts utilizing both thermal and non-thermal effects with the temperature gradient in the catalyst bed. Adapted from ref. [98] with permission. Copyright 2018 American Chemical Society. (C) Al@MIL-53 for enlarged surface area for gas uptake and enhanced reverse water-gas shift reaction. Adapted from ref. [99] with permission. Copyright 2019 The American Association for the advancement of science. (D) Schematic of hot-electron induced chemistry at Al-Cu2O interface for reverse water-gas shift reaction. Adapted from ref. [5] with permission. Copyright 2017 Springer Nature. (E) In-operando FTIR measurement of reaction steps under different temperatures for photothermal CO2 methanation on a Rh/al catalyst. Adapted from ref. [8] with permission. Copyright 2021 American Chemical Society.
Summary of gas-phase photocatalytic CO2 hydrogenation.
Nanostructure | Light Source | Power density | Selectivity | The highlight | Reference |
---|---|---|---|---|---|
of the work | number | ||||
Rh/Al2O3 | UV (365 nm) LED | 3 W/cm2 | >98 % to methane | Activation energy decrease and selectivity control to methanation | [9] |
Blue (400 nm) LED | >86 % to methane | ||||
Rh/TiO2 | UV (365 nm) LED | 2.24 W/cm2 | >98 % to methane in both light and dark | 46 % quantum efficiency at 350 °C surface temperature | [98] |
Blue (400 nm) LED | 3.55 W/cm2 | ||||
Al@Cu2O | Supercontinuum laser | 10 W/cm2 | >99.3 % toward CO | 3 % quantum efficiency. Hot electrons in the Cu2O shell drive the transformation | [5] |
Al@MOF | Supercontinuum laser | 350 mW | CO | Enhanced gas uptake increases the reactivity | [99] |
Rh/Al | Solar simulator | 11.3 W/cm2 | Nearly 100 % selectivity towards methane. | CH4 production rate at 550 mmol g−1 h−1 under a concentrated solar simulated source (11.3 W/cm2) irradiation. And In-operando FTIR measurement for multiple steps activation and conversion. | [8] |
4 Electromagnetic and quantum predictions of photochemical transformations
4.1 Wavelength-dependent photochemical transformations
For intermediate-sized nanoparticles (>5 nm and <100 nm), Mie theory and Gans theory are often employed to model the dielectric constant of a metal to analytically determine the absorption, scattering, and extinction spectra of metal nanoparticles with varying shapes, sizes, compositions, and dielectric environments. These methods are described in detail in several excellent review papers [28, 100]. Numerical methods such as finite-difference time-domain (FDTD), boundary element method (BEM), and finite element methods (FEM), can also be used to simulate and calculate the optical properties of the plasmonic nanostructure with high accuracy by solving Maxwell’s equations in three dimensions, depending on the dimension, geometry, symmetry, etc. In addition to simulating spectra and correlating these with experimental extinction, diffuse reflectance, dark-field scattering, etc., classical or semi-classical electromagnetic calculations offer us a tool to distinguish the mechanism of various reaction pathways and to interpret wavelength-dependent reactivity of plasmonic photocatalysts. An early example of utilizing this tool to distinguish the hot electrons in metal includes Zheng, B. et al., in which they fabricated a Au-TiO2 device and measured the photocurrent with the Schottky contact and Ohmic contact between Au and TiO2 interface [101]. The authors exploit the fact that only the plasmon-induced hot electrons have sufficient energy to overcome the Schottky barrier. The study confirmed that hot carrier generation within a plasmonic metal is directly correlated to the volume integration of |E|2 within the electron mean free path of the metal. The same idea can also be applied to Al@Cu2O core–shell nanoparticles for reverse water-gas shift, Al-Pd antenna-reactors for H2 dissociation and C–F activation, and solvated electrons from Au and Ag nanoparticles [5, 102–104]. There are many publications which correlate absorption or extinction cross-sections with wavelength-dependent reactivity, especially when the classic absorption cross-section is very close to its extinction cross-section in Au nanoparticles [31, 53, 58, 62]. However, the portion of the energy or charge transfer process directly taking part in the chemistry is not clear in this paradigm. Yuan., L., et al. developed a model based on the semi-classical hot electron theory applied to aluminum plasmonic nanocrystals from a previous report [34, 105]. The interband hot electrons were accounted for by including the contribution of the interband transition to the imaginary part of the permittivity in the |E|2 integration. Besides the peak originating from the plasmon modes, only the hot-electron cross-section follows the increasing trend at 700–800 nm, which matches well with the experiment. The strong quantum confinement due to the small size or features of nanoparticles also give rise to phonon-assisted Landau damping (intraband transition) hot carriers. These are usually considered a bonus to the classic absorption cross-section, but one can consider adding the corresponding term into the dielectric constant of the metal just like the Drude electron term and Lorentzian oscillators for interband transition (or another model such as Brendal–Bormann, and quantum mechanical model, etc.). [106, 107]. The semi-classical theory can easily decouple the hot electron cross-section from direct photoexcitation (interband transition), Landau damping (intraband transition), and photothermal effects by comparing the spectra with the wavelength-dependent reactivity or quantum efficiency.
Numerous challenges remain in electromagnetic calculations, such as taking into account the detailed picture of the charge transfer of hot electrons to the anti-bonding orbitals of the molecule (or hot holes to bonding orbitals). An illustrative example is the utilization of a Lorentzian function in hot carrier integration. This function effectively characterizes the energy position distribution of a molecular adsorbate, enabling hot carriers produced at specific energy levels to be injected into the adsorbates, consequently prompting the desired chemical reactions [79, 108]. Whether the reaction rate or external quantum efficiency should be chosen to correlate the calculated hot electron cross-section (or rate) depends on how the hot carriers impacts the conversion of molecules and whether it is one hot electron corresponding to one molecule or if carrier multiplication also contributes to the reaction [85, 109]. Moving forward, further developments in high-accuracy quantum chemistry calculations, ultrafast pump-probe measurements, and in-operando measurements of reaction intermediates will be needed to obtain accurate information on these processes.
5 Quantum mechanistic insights on plasmonic photocatalysis
To better understand the mechanisms underlying plasmon catalysis, considerable effort has been devoted to computationally predicting excited-state processes. Such calculations are necessary to explain plasmon creation and dephasing, as well as the excited-state photochemistry arising from the presence of hot carriers and non-equilibrium phonon distributions.
To describe light–matter interaction in nanoparticles, first-principles atomistic calculations based on density functional theory (DFT) and time-dependent DFT (TDDFT) can only simulate systems of relatively small sizes (typically nanoparticles of hundreds or few thousands of atoms). Still, they offer unique quantum-mechanical insights into the physical origin of plasmon dephasing and charge transfer [110–118]. The other extreme limit is that of a nanoparticle large enough that its plasmon dephasing and hot electron generation can be modeled from a simpler geometry by considering a semi-infinite metallic slab. Accurate computations were imperative in elucidating the contribution of d bands in noble metals to the production of hot electrons from surface plasmon polaritons [119, 120], as well as the impact of phonons and finite-size effects on plasmon damping [22].
A major challenge in predicting excited-state chemistry is to accurately address the interplay between local chemical environment and excited-state electronic and optical properties. Quantum mechanical, atomistic effects, and electron correlation beyond what is available with mean-field formalisms such as DFT must be included. There are a few directions that the community is exploring towards this goal. First, there has been qualitative success in the utilization of embedding techniques that treat chemically active sites with a high-level quantum chemistry theory, while the remainder of the nanoparticle (or approximate planar slab geometry) is treated approximately within DFT. For instance, embedding theory was recently used to elucidate the hot-electron-based splitting of H2 on an Au surface [121].
There are also numerous examples employing high-accuracy excited-state calculations to identify the potential transient negative ion (TNI) state to explain the reduction of energy barriers for key intermediate steps compared to a ground-state reaction pathway, helping to elucidate the unique properties of plasmon-induced chemistry [31, 108, 121].
In addition, theoretical approaches based on quantum electron dynamics (QED), wherein one explicitly includes the quantized photon modes together with interacting electrons treated within DFT or correlated wavefunction approaches, allows one to include strong electromagnetic fields when describing such chemical reactions [122]. Finally, emerging large-scale quantum many-body formalisms [123–126], such as those based on the GW approximation to the electronic self-energy [127] and the Bethe-Salpeter equation (BSE) [128], are poised to address (i) the coupling of light with plasmons and other possible excited states in the nanoparticles; (ii) the atomistic details of the reaction, including realistic surface reconstruction, dopant distributions, and adsorbate reaction coordinates; and (iii) the net results of these excitations into the overall chemical reaction, including any necessary quantum many-body interactions.
6 Summary and outlook
In this article, we introduced the unique features of plasmonic photocatalysis and potential opportunities to use renewable light energy to drive sustainable chemistry. We explored recent progress in three promising applications: plasmon-induced ammonia synthesis, gas-phase hydrogen generation, and CO2 reduction. The importance of these three applications cannot be understated, as they are critical for the sustainable production of food, fuel, and feedstock for organic/pharmaceutical synthesis. Electromagnetic theory and quantum chemistry are two toolboxes which provide unique insight into light–matter interactions and surface-adsorbate interfaces. Experiment and theory are both necessary to determine the underlying mechanisms in plasmonic photocatalysis and for further optimization of the light–matter and molecule-active center interactions to synergistically increase the yield and maximize the selectivity for future applications and industrialization.
Future work in the field will include more advanced in-situ and in-operando measurements such as employing optically-coupled environmental transmission electron microscopy to understand light-induced phase transformations beyond hydrides [129–131] on the sub-nm scale. FTIR and Raman spectroscopy can also be employed to identify key intermediates on the surface of catalysts. High-accuracy quantum mechanical simulation methods for both light–matter interactions and excited-state reaction pathways are also expected to be developed. Resolving the time-scale of critical intermediate species and plasmon lifetime on different surfaces remains a significant challenge. The field should focus on obtaining a detailed picture of the photochemistry, especially for small molecules with fewer side reactions, through quantitative measurements. Subsequently, researchers need to probe and control the more complicated chemistry, such as the Fischer–Tropsch synthesis, by understanding the multiphoton and hot-carrier processes. Looking forward, the use of earth-abundant materials for catalysts, renewable light sources such as solar light and LED illumination, recycling of noble metal catalysts, and scaling reactor sizes are all critical developments toward the goal of sustainable chemistry. The use of green energy sources in the chemical industry will be vital. Taking the steel industry as an example, both using green hydrogen for CO2 reduction and developing composite materials with the necessary mechanical and corrosion properties would significantly reduce the industries carbon footprint and support the development of sustainable chemistry.
Funding source: National Science Foundation Graduate Research Fellowship
Award Identifier / Grant number: DGE-1656518
Funding source: W. M. Keck Foundation
Award Identifier / Grant number: 994816
-
Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
-
Research funding: All authors acknowledge the support from Keck Foundation under grant No. 994816. B.B.B. was supported by the National Science Foundation Graduate Research Fellowship under grant number DGE-1656518.
-
Conflict of interest statement: The authors declare no conflicts of interest regarding this article.
References
[1] J. R. Ross, Heterogeneous Catalysis: Fundamentals and Applications, Amsterdam, Elsevier, 2012.Search in Google Scholar
[2] S. Linic, P. Christopher, and D. B. Ingram, “Plasmonic-metal nanostructures for efficient conversion of solar to chemical energy,” Nat. Mater., vol. 10, no. 12, pp. 911–921, 2011. https://doi.org/10.1038/nmat3151.Search in Google Scholar PubMed
[3] A. Marimuthu, J. Zhang, and S. Linic, “Tuning selectivity in propylene epoxidation by plasmon mediated photo-switching of Cu oxidation state,” Science, vol. 339, no. 6127, pp. 1590–1593, 2013. https://doi.org/10.1126/science.1231631.Search in Google Scholar PubMed
[4] D. F. Swearer, H. Zhao, L. Zhou, et al.., “Heterometallic antenna-reactor complexes for photocatalysis,” Proc. Natl. Acad. Sci. U. S. A., vol. 113, no. 32, pp. 8916–8920, 2016. https://doi.org/10.1073/pnas.1609769113.Search in Google Scholar PubMed PubMed Central
[5] H. Robatjazi, H. Zhao, D. F. Swearer, et al.., “Plasmon-induced selective carbon dioxide conversion on earth-abundant aluminum-cuprous oxide antenna-reactor nanoparticles,” Nat. Commun., vol. 8, no. 1, p. 27, 2017. https://doi.org/10.1038/s41467-017-00055-z.Search in Google Scholar PubMed PubMed Central
[6] G. Kumari, X. Zhang, D. Devasia, J. Heo, and P. K. Jain, “Watching visible light-driven CO2 reduction on a plasmonic nanoparticle catalyst,” ACS Nano, vol. 12, no. 8, pp. 8330–8340, 2018. https://doi.org/10.1021/acsnano.8b03617.Search in Google Scholar PubMed
[7] S. Yu, A. J. Wilson, J. Heo, and P. K. Jain, “Plasmonic control of multi-electron transfer and C-C coupling in visible-light-driven CO2 reduction on Au nanoparticles,” Nano Lett., vol. 18, no. 4, pp. 2189–2194, 2018. https://doi.org/10.1021/acs.nanolett.7b05410.Search in Google Scholar PubMed
[8] G. Fu, M. Jiang, J. Liu, et al.., “Rh/Al nanoantenna photothermal catalyst for wide-spectrum solar-driven CO2 methanation with nearly 100% selectivity,” Nano Lett., vol. 21, no. 20, pp. 8824–8830, 2021. https://doi.org/10.1021/acs.nanolett.1c03215.Search in Google Scholar PubMed
[9] X. Zhang, X. Li, D. Zhang, et al.., “Product selectivity in plasmonic photocatalysis for carbon dioxide hydrogenation,” Nat. Commun., vol. 8, no. 1, p. 14542, 2017. https://doi.org/10.1038/ncomms14542.Search in Google Scholar PubMed PubMed Central
[10] S. Luo, H. Lin, Q. Wang, et al.., “Triggering water and methanol activation for solar-driven H2 production: interplay of dual active sites over plasmonic ZnCu alloy,” J. Am. Chem. Soc., vol. 143, no. 31, pp. 12145–12153, 2021. https://doi.org/10.1021/jacs.1c04315.Search in Google Scholar PubMed
[11] Z. Li, J. Liu, J. Zhao, et al.., “Photo-driven hydrogen production from methanol and water using plasmonic Cu nanoparticles derived from layered double hydroxides,” Adv. Funct. Mater., vol. 33, no. 11, p. 2213672, 2022. https://doi.org/10.1002/adfm.202213672.Search in Google Scholar
[12] E. Cortés, L. V. Besteiro, A. Alabastri, et al.., “Challenges in plasmonic catalysis,” ACS Nano, vol. 14, no. 12, pp. 16202–16219, 2020. https://doi.org/10.1021/acsnano.0c08773.Search in Google Scholar PubMed
[13] Y. Xin, K. Yu, L. Zhang, et al.., “Copper-based plasmonic catalysis: recent advances and future perspectives,” Adv. Mater., vol. 33, no. 32, p. e2008145, 2021. https://doi.org/10.1002/adma.202008145.Search in Google Scholar PubMed
[14] B. Puértolas, M. Comesaña-Hermo, L. V. Besteiro, M. Vázquez-González, and M. A. Correa-Duarte, “Challenges and opportunities for renewable ammonia production via plasmon-assisted photocatalysis,” Adv. Energy Mater., vol. 12, no. 18, p. 2103909, 2022. https://doi.org/10.1002/aenm.202103909.Search in Google Scholar
[15] S. Ezendam, M. Herran, L. Nan, et al.., “Hybrid plasmonic nanomaterials for hydrogen generation and carbon dioxide reduction,” Adv. Funct. Mater., vol. 7, no. 2, pp. 778–815, 2022. https://doi.org/10.1021/acsenergylett.1c02241.Search in Google Scholar PubMed PubMed Central
[16] W. Jiang, B. Q. L. Low, R. Long, et al.., “Active site engineering on plasmonic nanostructures for efficient photocatalysis,” ACS Nano, vol. 17, no. 5, pp. 4193–4229, 2023. https://doi.org/10.1021/acsnano.2c12314.Search in Google Scholar PubMed
[17] C. E. Finke, H. F. Leandri, E. T. Karumb, D. Zheng, M. R. Hoffmann, and N. A. Fromer, “Economically advantageous pathways for reducing greenhouse gas emissions from industrial hydrogen under common, current economic conditions,” Energy Environ. Sci., vol. 14, no. 3, pp. 1517–1529, 2021. https://doi.org/10.1039/D0EE03768K.Search in Google Scholar
[18] I. E. Agency, CO2 Emissions in 2022. Available at: https://www.iea.org/reports/co2-emissions-in-2022 [accessed: Apr. 16, 2023].Search in Google Scholar
[19] U. Aslam, V. G. Rao, S. Chavez, and S. Linic, “Catalytic conversion of solar to chemical energy on plasmonic metal nanostructures,” Nat. Catal., vol. 1, no. 9, pp. 656–665, 2018. https://doi.org/10.1038/s41929-018-0138-x.Search in Google Scholar
[20] N. J. Halas, “Spiers Memorial Lecture Introductory lecture: hot-electron science and microscopic processes in plasmonics and catalysis,” Faraday Discuss., vol. 214, pp. 13–33, 2019. https://doi.org/10.1039/C9FD00001A.Search in Google Scholar
[21] M. L. Brongersma, N. J. Halas, and P. Nordlander, “Plasmon-induced hot carrier science and technology,” Nat. Nanotechnol., vol. 10, no. 1, pp. 25–34, 2015. https://doi.org/10.1038/nnano.2014.311.Search in Google Scholar PubMed
[22] A. M. Brown, R. Sundararaman, P. Narang, W. A. Goddard3rd, and H. A. Atwater, “Nonradiative plasmon decay and hot carrier dynamics: effects of phonons, surfaces, and geometry,” ACS Nano, vol. 10, no. 1, pp. 957–966, 2016. https://doi.org/10.1021/acsnano.5b06199.Search in Google Scholar PubMed
[23] A. O. Govorov, H. Zhang, and Y. K. Gun’ko, “Theory of photoinjection of hot plasmonic carriers from metal nanostructures into semiconductors and surface molecules,” J. Phys. Chem. C, vol. 117, no. 32, pp. 16616–16631, 2013. https://doi.org/10.1021/jp405430m.Search in Google Scholar
[24] L. Chang, L. V. Besteiro, J. Sun, et al.., “Electronic structure of the plasmons in metal nanocrystals: fundamental limitations for the energy efficiency of hot electron generation,” ACS Energy Lett., vol. 4, no. 10, pp. 2552–2568, 2019. https://doi.org/10.1021/acsenergylett.9b01617.Search in Google Scholar
[25] S. Linic, U. Aslam, C. Boerigter, and M. Morabito, “Photochemical transformations on plasmonic metal nanoparticles,” Nat. Mater., vol. 14, no. 6, pp. 567–576, 2015. https://doi.org/10.1038/nmat4281.Search in Google Scholar PubMed
[26] M. W. Knight, N. S. King, L. Liu, H. O. Everitt, P. Nordlander, and N. J. Halas, “Aluminum for plasmonics,” ACS Nano, vol. 8, no. 1, pp. 834–840, 2014. https://doi.org/10.1021/nn405495q.Search in Google Scholar PubMed
[27] D. F. Swearer, R. K. Leary, R. Newell, et al.., “Transition-metal decorated aluminum nanocrystals,” ACS Nano, vol. 11, no. 10, pp. 10281–10288, 2017. https://doi.org/10.1021/acsnano.7b04960.Search in Google Scholar PubMed
[28] K. M. Mayer and J. H. Hafner, “Localized surface plasmon resonance sensors,” Chem. Rev., vol. 111, no. 6, pp. 3828–3857, 2011. https://doi.org/10.1021/cr100313v.Search in Google Scholar PubMed
[29] B. Seemala, A. J. Therrien, M. Lou, et al.., “Plasmon-mediated catalytic O2 dissociation on Ag nanostructures: hot electrons or near fields?” ACS Energy Lett., vol. 4, no. 8, pp. 1803–1809, 2019. https://doi.org/10.1021/acsenergylett.9b00990.Search in Google Scholar
[30] Y. Kim, D. Dumett Torres, and P. K. Jain, “Activation energies of plasmonic catalysts,” Nano Lett., vol. 16, no. 5, pp. 3399–3407, 2016. https://doi.org/10.1021/acs.nanolett.6b01373.Search in Google Scholar PubMed
[31] M. Lou, J. L. Bao, L. Zhou, et al.., “Direct H2S decomposition by plasmonic photocatalysis: efficient remediation plus sustainable hydrogen production,” ACS Energy Lett., vol. 7, no. 10, pp. 3666–3674, 2022. https://doi.org/10.1021/acsenergylett.2c01755.Search in Google Scholar
[32] E. L. Keller and R. R. Frontiera, “Ultrafast nanoscale Raman thermometry proves heating is not a primary mechanism for plasmon-driven photocatalysis,” ACS Nano, vol. 12, no. 6, pp. 5848–5855, 2018. https://doi.org/10.1021/acsnano.8b01809.Search in Google Scholar PubMed
[33] L. Zhou, C. Zhang, M. J. McClain, et al.., “Aluminum nanocrystals as a plasmonic photocatalyst for hydrogen dissociation,” Nano Lett., vol. 16, no. 2, pp. 1478–1484, 2016. https://doi.org/10.1021/acs.nanolett.5b05149.Search in Google Scholar PubMed
[34] L. Yuan, M. Lou, B. D. Clark, et al.., “Morphology-dependent reactivity of a plasmonic photocatalyst,” ACS Nano, vol. 14, no. 9, pp. 12054–12063, 2020. https://doi.org/10.1021/acsnano.0c05383.Search in Google Scholar PubMed
[35] A. Al-Zubeidi, B. S. Hoener, S. S. E. Collins, et al.., “Hot holes assist plasmonic nanoelectrode dissolution,” Nano Lett., vol. 19, no. 2, pp. 1301–1306, 2019. https://doi.org/10.1021/acs.nanolett.8b04894.Search in Google Scholar PubMed
[36] J. S. DuChene, G. Tagliabue, A. J. Welch, W. H. Cheng, and H. A. Atwater, “Hot hole collection and photoelectrochemical CO2 reduction with plasmonic Au/p-GaN photocathodes,” Nano Lett., vol. 18, no. 4, pp. 2545–2550, 2018. https://doi.org/10.1021/acs.nanolett.8b00241.Search in Google Scholar PubMed
[37] J. S. DuChene, G. Tagliabue, A. J. Welch, X. Li, W. H. Cheng, and H. A. Atwater, “Optical excitation of a nanoparticle Cu/p-NiO photocathode improves reaction selectivity for CO2 reduction in aqueous electrolytes,” Nano Lett., vol. 20, no. 4, pp. 2348–2358, 2020. https://doi.org/10.1021/acs.nanolett.9b04895.Search in Google Scholar PubMed
[38] Y. Y. Cai, S. S. E. Collins, M. J. Gallagher, et al.., “Single-particle emission spectroscopy resolves d-hole relaxation in copper nanocubes,” ACS Energy Lett., vol. 4, no. 10, pp. 2458–2465, 2019. https://doi.org/10.1021/acsenergylett.9b01747.Search in Google Scholar
[39] G. Tagliabue, J. S. DuChene, M. Abdellah, et al.., “Ultrafast hot-hole injection modifies hot-electron dynamics in Au/p-GaN heterostructures,” Nat. Mater., vol. 19, no. 12, pp. 1312–1318, 2020. https://doi.org/10.1038/s41563-020-0737-1.Search in Google Scholar PubMed
[40] L. Wang, M. Hasanzadeh Kafshgari, and M. Meunier, “Optical properties and applications of plasmonic‐metal nanoparticles,” Adv. Funct. Mater., vol. 30, no. 51, p. 2005400, 2020. https://doi.org/10.1002/adfm.202005400.Search in Google Scholar
[41] T. C. Asmara, D. Wan, Y. Zhao, et al.., “Tunable and low-loss correlated plasmons in Mott-like insulating oxides,” Nat. Commun., vol. 8, no. 1, p. 15271, 2017. https://doi.org/10.1038/ncomms15271.Search in Google Scholar PubMed PubMed Central
[42] D. Wan, B. Yan, J. Chen, et al.., “New family of plasmonic photocatalysts without noble metals,” Chem. Mater., vol. 31, no. 7, pp. 2320–2327, 2019. https://doi.org/10.1021/acs.chemmater.8b04185.Search in Google Scholar
[43] D. F. Swearer, N. R. Knowles, H. O. Everitt, and N. J. Halas, “Light-driven chemical looping for ammonia synthesis,” ACS Energy Lett., vol. 4, no. 7, pp. 1505–1512, 2019. https://doi.org/10.1021/acsenergylett.9b00860.Search in Google Scholar
[44] H. Cheng, Y. Kuwahara, and H. Yamashita, “Earth-abundant plasmonic catalysts,” in Plasmonic Catalysis: From Fundamentals to Applications, vol. 10, P. H. C. Camargo and E. Cortés, Eds., Weinheim, Wiley, 2021, p. 3.Search in Google Scholar
[45] J. W. Erisman, M. A. Sutton, J. Galloway, Z. Klimont, and W. Winiwarter, “How a century of ammonia synthesis changed the world,” Nat. Geosci., vol. 1, no. 10, pp. 636–639, 2008. https://doi.org/10.1038/ngeo325.Search in Google Scholar
[46] C. Smith, A. K. Hill, and L. Torrente-Murciano, “Current and future role of Haber–Bosch ammonia in a carbon-free energy landscape,” Energy Environ. Sci., vol. 13, no. 2, pp. 331–344, 2020. https://doi.org/10.1039/C9EE02873K.Search in Google Scholar
[47] P. Mehta, P. Barboun, F. A. Herrera, et al.., “Overcoming ammonia synthesis scaling relations with plasma-enabled catalysis,” Nat. Catal., vol. 1, no. 4, pp. 269–275, 2018. https://doi.org/10.1038/s41929-018-0045-1.Search in Google Scholar
[48] C. Zhang, Z. Wang, J. Lei, L. Ma, B. I. Yakobson, and J. M. Tour, “Atomic molybdenum for synthesis of ammonia with 50% faradic efficiency,” Small, vol. 18, no. 15, p. e2106327, 2022. https://doi.org/10.1002/smll.202106327.Search in Google Scholar PubMed
[49] X. Liu, A. Elgowainy, and M. Wang, “Life cycle energy use and greenhouse gas emissions of ammonia production from renewable resources and industrial by-products,” Green Chem., vol. 22, no. 17, pp. 5751–5761, 2020. https://doi.org/10.1039/D0GC02301A.Search in Google Scholar
[50] B. Jabarivelisdeh, E. Jin, P. Christopher, and E. Masanet, “Model-based analysis of ammonia production processes for quantifying energy use, emissions, and reduction potentials,” ACS Sustainable Chem. Eng., vol. 10, no. 49, pp. 16280–16289, 2022. https://doi.org/10.1021/acssuschemeng.2c04976.Search in Google Scholar
[51] B.-Y. Zhang, H.-Y. Su, J.-X. Liu, and W.-X. Li, “Interplay between bite activity and density of BCC iron for ammonia synthesis based on first‐principles theory,” ChemCatChem, vol. 11, no. 7, pp. 1928–1934, 2019.10.1002/cctc.201900175Search in Google Scholar
[52] H. Zeng, S. Terazono, and T. Tanuma, “A novel catalyst for ammonia synthesis at ambient temperature and pressure: visible light responsive photocatalyst using localized surface plasmon resonance,” Catal. Commun., vol. 59, pp. 40–44, 2015. https://doi.org/10.1016/j.catcom.2014.09.034.Search in Google Scholar
[53] C. Hu, X. Chen, J. Jin, et al.., “Surface plasmon enabling nitrogen fixation in pure water through a dissociative mechanism under mild conditions,” J. Am. Chem. Soc., vol. 141, no. 19, pp. 7807–7814, 2019. https://doi.org/10.1021/jacs.9b01375.Search in Google Scholar PubMed
[54] J. Yang, Y. Guo, R. Jiang, et al.., “High-efficiency "Working-in-Tandem" nitrogen photofixation achieved by assembling plasmonic gold nanocrystals on ultrathin titania nanosheets,” J. Am. Chem. Soc., vol. 140, no. 27, pp. 8497–8508, 2018. https://doi.org/10.1021/jacs.8b03537.Search in Google Scholar PubMed
[55] H. Jia, A. Du, H. Zhang, et al.., “Site-selective growth of crystalline ceria with oxygen vacancies on gold nanocrystals for near-infrared nitrogen photofixation,” J. Am. Chem. Soc., vol. 141, no. 13, pp. 5083–5086, 2019. https://doi.org/10.1021/jacs.8b13062.Search in Google Scholar PubMed
[56] S. Wu, Z. Chen, K. Liu, W. Yue, L. Wang, and J. Zhang, “Chemisorption-induced and plasmon-promoted photofixation of nitrogen on gold-loaded carbon nitride nanosheets,” ChemSusChem, vol. 13, no. 13, pp. 3455–3461, 2020. https://doi.org/10.1002/cssc.202000818.Search in Google Scholar PubMed
[57] Y. Guo, J. Yang, D. Wu, et al.., “Au nanoparticle-embedded, nitrogen-deficient hollow mesoporous carbon nitride spheres for nitrogen photofixation,” J. Mater. Chem. A, vol. 8, no. 32, pp. 16218–16231, 2020. https://doi.org/10.1039/d0ta03793a.Search in Google Scholar
[58] L. W. Chen, Y. C. Hao, Y. Guo, et al.., “Metal-organic framework membranes encapsulating gold nanoparticles for direct plasmonic photocatalytic nitrogen fixation,” J. Am. Chem. Soc., vol. 143, no. 15, pp. 5727–5736, 2021. https://doi.org/10.1021/jacs.0c13342.Search in Google Scholar PubMed
[59] X. Li, X. Zhang, H. O. Everitt, and J. Liu, “Light-induced thermal gradients in ruthenium catalysts significantly enhance ammonia production,” Nano Lett., vol. 19, no. 3, pp. 1706–1711, 2019. https://doi.org/10.1021/acs.nanolett.8b04706.Search in Google Scholar PubMed
[60] C. Mao, L. Yu, J. Li, J. Zhao, and L. Zhang, “Energy-confined solar thermal ammonia synthesis with K/Ru/TiO2-xHx,” Appl. Catal., B, vol. 224, pp. 612–620, 2018. https://doi.org/10.1016/j.apcatb.2017.11.010.Search in Google Scholar
[61] C. Mao, H. Li, H. Gu, et al.., “Beyond the thermal equilibrium limit of ammonia synthesis with dual temperature zone catalyst powered by solar light,” Chem, vol. 5, no. 10, pp. 2702–2717, 2019. https://doi.org/10.1016/j.chempr.2019.07.021.Search in Google Scholar
[62] T. Hou, L. Chen, Y. Xin, et al.., “Porous CuFe for plasmon-assisted N2 photofixation,” ACS Energy Lett., vol. 5, no. 7, pp. 2444–2451, 2020. https://doi.org/10.1021/acsenergylett.0c00959.Search in Google Scholar
[63] L. Chen, Y. Wang, F. Yu, X. Shen, and C. Duan, “A simple strategy for engineering heterostructures of Au nanoparticle-loaded metal–organic framework nanosheets to achieve plasmon-enhanced photocatalytic CO2 conversion under visible light,” J. Mater. Chem. A, vol. 7, no. 18, pp. 11355–11361, 2019. https://doi.org/10.1039/c9ta01840a.Search in Google Scholar
[64] Y. Manoharan, S. E. Hosseini, B. Butler, et al.., “Hydrogen fuel cell vehicles; current status and future prospect,” Appl. Sci., vol. 9, no. 11, pp. 1–17, 2019. https://doi.org/10.3390/app9112296.Search in Google Scholar
[65] J. Jia, L. C. Seitz, J. D. Benck, et al.., “Solar water splitting by photovoltaic-electrolysis with a solar-to-hydrogen efficiency over 30%,” Nat. Commun., vol. 7, no. 1, p. 13237, 2016. https://doi.org/10.1038/ncomms13237.Search in Google Scholar PubMed PubMed Central
[66] E. Santos, P. Quaino, and W. Schmickler, “Theory of electrocatalysis: hydrogen evolution and more,” Phys. Chem. Chem. Phys., vol. 14, no. 32, pp. 11224–11233, 2012. https://doi.org/10.1039/c2cp40717e.Search in Google Scholar PubMed
[67] K. Maeda, “Photocatalytic water splitting using semiconductor particles: history and recent developments,” J. Photochem. Photobiol. C: Photochem. Rev., vol. 12, no. 4, pp. 237–268, 2011. https://doi.org/10.1016/j.jphotochemrev.2011.07.001.Search in Google Scholar
[68] L. Yuan, A. Kuriakose, J. Zhou, H. Robatjazi, P. Nordlander, and N. J. Halas, “Plasmonically enhanced hydrogen evolution with an Al–TiO2-Based photoelectrode,” J. Phys. Chem. C, vol. 126, no. 32, pp. 13714–13719, 2022. https://doi.org/10.1021/acs.jpcc.2c03961.Search in Google Scholar
[69] A. Fujishima and K. Honda, “Electrochemical evidence for the mechanism of the primary stage of photosynthesis,” Bull. Chem. Soc. Jpn., vol. 44, no. 4, pp. 1148–1150, 1971. https://doi.org/10.1246/bcsj.44.1148.Search in Google Scholar
[70] R. Marschall, “50 years of materials research for photocatalytic water splitting,” Eur. J. Inorg., vol. 2021, no. 25, pp. 2435–2441, 2021. https://doi.org/10.1002/ejic.202100264.Search in Google Scholar
[71] S. Mubeen, J. Lee, N. Singh, S. Krämer, G. D. Stucky, and M. Moskovits, “An autonomous photosynthetic device in which all charge carriers derive from surface plasmons,” Nat. Nanotechnol., vol. 8, no. 4, pp. 247–251, 2013. https://doi.org/10.1038/nnano.2013.18.Search in Google Scholar PubMed
[72] G. Graziano, “All-plasmonic water splitting,” Nat. Nanotechnol., vol. 16, no. 10, p. 1053, 2021. https://doi.org/10.1038/s41565-021-00991-4.Search in Google Scholar PubMed
[73] H. Robatjazi, S. M. Bahauddin, C. Doiron, and I. Thomann, “Direct plasmon-driven photoelectrocatalysis,” Nano Lett., vol. 15, no. 9, pp. 6155–6161, 2015. https://doi.org/10.1021/acs.nanolett.5b02453.Search in Google Scholar PubMed
[74] K. Song, H. Lee, M. Lee, and J. Y. Park, “Plasmonic hot hole-driven water splitting on Au nanoprisms/P-type GaN,” ACS Energy Lett., vol. 6, no. 4, pp. 1333–1339, 2021. https://doi.org/10.1021/acsenergylett.1c00366.Search in Google Scholar
[75] X. Shi, K. Ueno, T. Oshikiri, Q. Sun, K. Sasaki, and H. Misawa, “Enhanced water splitting under modal strong coupling conditions,” Nat. Nanotechnol., vol. 13, no. 10, pp. 953–958, 2018. https://doi.org/10.1038/s41565-018-0208-x.Search in Google Scholar PubMed
[76] P. Subramanyam, B. Meena, V. Biju, H. Misawa, and S. Challapalli, “Emerging materials for plasmon-assisted photoelectrochemical water splitting,” J. Photochem. Photobiol. C: Photochem. Rev., vol. 51, p. 100472, 2022. https://doi.org/10.1016/j.jphotochemrev.2021.100472.Search in Google Scholar
[77] F. Dawood, M. Anda, and G. M. Shafiullah, “Hydrogen production for energy: an overview,” Int. J. Hydrog. Energy, vol. 45, no. 7, pp. 3847–3869, 2020. https://doi.org/10.1016/j.ijhydene.2019.12.059.Search in Google Scholar
[78] L. Zhou, D. F. Swearer, C. Zhang, et al.., “Quantifying hot carrier and thermal contributions in plasmonic photocatalysis,” Science, vol. 362, no. 6410, pp. 69–72, 2018. https://doi.org/10.1126/science.aat6967.Search in Google Scholar PubMed
[79] Y. Yuan, L. Zhou, H. Robatjazi, et al.., “Earth-abundant photocatalyst for H2 generation from NH3 with light-emitting diode illumination,” Science, vol. 378, no. 6622, pp. 889–893, 2022. https://doi.org/10.1126/science.abn5636.Search in Google Scholar PubMed
[80] L. Zhou, J. M. P. Martirez, J. Finzel, et al.., “Light-driven methane dry reforming with single atomic site antenna-reactor plasmonic photocatalysts,” Nat. Energy, vol. 5, no. 1, pp. 61–70, 2020. https://doi.org/10.1038/s41560-019-0517-9.Search in Google Scholar
[81] Z.-Y. Zhang, T. Zhang, R.-K. Wang, et al.., “Photo-enhanced dry reforming of methane over Pt-Au/P25 composite catalyst by coupling plasmonic effect,” J. Catal., vol. 413, pp. 829–842, 2022. https://doi.org/10.1016/j.jcat.2022.07.028.Search in Google Scholar
[82] B. L. Farrell, V. O. Igenegbai, and S. Linic, “A viewpoint on direct methane conversion to ethane and ethylene using oxidative coupling on solid catalysts,” ACS Catal., vol. 6, no. 7, pp. 4340–4346, 2016. https://doi.org/10.1021/acscatal.6b01087.Search in Google Scholar
[83] D. Takami, Y. Ito, S. Kawaharasaki, A. Yamamoto, and H. Yoshida, “Low temperature dry reforming of methane over plasmonic Ni photocatalysts under visible light irradiation,” Sustain. Energy Fuels, vol. 3, no. 11, pp. 2968–2971, 2019. https://doi.org/10.1039/c9se00206e.Search in Google Scholar
[84] A. R. Woldu, Z. Huang, P. Zhao, L. Hu, and D. Astruc, “Electrochemical CO2 reduction (CO2RR) to multi-carbon products over copper-based catalysts,” Coord. Chem. Rev., vol. 454, p. 214340, 2022. https://doi.org/10.1016/j.ccr.2021.214340.Search in Google Scholar
[85] L. Zhou, M. Lou, J. L. Bao, et al.., “Hot carrier multiplication in plasmonic photocatalysis,” Proc. Natl. Acad. Sci. U. S. A., vol. 118, no. 20, 2021, Art. no. e2022109118. https://doi.org/10.1073/pnas.2022109118.Search in Google Scholar PubMed PubMed Central
[86] C. Zhan, Q. X. Wang, J. Yi, et al.., “Plasmonic nanoreactors regulating selective oxidation by energetic electrons and nanoconfined thermal fields,” Sci. Adv., vol. 7, no. 10, p. eabf0962, 2021. https://doi.org/10.1126/sciadv.abf0962.Search in Google Scholar PubMed PubMed Central
[87] P. D. Dongare, Y. Zhao, D. Renard, et al.., “A 3D plasmonic antenna-reactor for nanoscale thermal hotspots and gradients,” ACS Nano, vol. 15, no. 5, pp. 8761–8769, 2021. https://doi.org/10.1021/acsnano.1c01046.Search in Google Scholar PubMed
[88] D. Devasia, A. J. Wilson, J. Heo, V. Mohan, and P. K. Jain, “A rich catalog of C-C bonded species formed in CO2 reduction on a plasmonic photocatalyst,” Nat. Commun., vol. 12, no. 1, p. 2612, 2021. https://doi.org/10.1038/s41467-021-22868-9.Search in Google Scholar PubMed PubMed Central
[89] W. Hou, W. H. Hung, P. Pavaskar, A. Goeppert, M. Aykol, and S. B. Cronin, “Photocatalytic conversion of CO2 to hydrocarbon fuels via plasmon-enhanced absorption and metallic interband transitions,” ACS Catal., vol. 1, no. 8, pp. 929–936, 2011. https://doi.org/10.1021/cs2001434.Search in Google Scholar
[90] J. Bian, Y. Qu, X. Zhang, N. Sun, D. Tang, and L. Jing, “Dimension-matched plasmonic Au/TiO2/BiVO4 nanocomposites as efficient wide-visible-light photocatalysts to convert CO2 and mechanistic insights,” J. Mater. Chem. A, vol. 6, no. 25, pp. 11838–11845, 2018. https://doi.org/10.1039/c8ta02889c.Search in Google Scholar
[91] K. M. Choi, D. Kim, B. Rungtaweevoranit, et al.., “Plasmon-enhanced photocatalytic CO2 conversion within metal-organic frameworks under visible light,” J. Am. Chem. Soc., vol. 139, no. 1, pp. 356–362, 2017. https://doi.org/10.1021/jacs.6b11027.Search in Google Scholar PubMed
[92] Y. Han, H. Xu, Y. Su, Z.-l. Xu, K. Wang, and W. Wang, “Noble metal (Pt, Au@Pd) nanoparticles supported on metal organic framework (MOF-74) nanoshuttles as high-selectivity CO2 conversion catalysts,” J. Catal., vol. 370, pp. 70–78, 2019. https://doi.org/10.1016/j.jcat.2018.12.005.Search in Google Scholar
[93] H. Zhang, T. Itoi, T. Konishi, and Y. Izumi, “Dual photocatalytic roles of light: charge separation at the band gap and heat via localized surface plasmon resonance to convert CO2 into CO over silver-zirconium oxide,” J. Am. Chem. Soc., vol. 141, no. 15, pp. 6292–6301, 2019. https://doi.org/10.1021/jacs.8b13894.Search in Google Scholar PubMed
[94] E. Liu, L. Qi, J. Bian, et al.., “A facile strategy to fabricate plasmonic Cu modified TiO2 nano-flower films for photocatalytic reduction of CO2 to methanol,” Mat. Res. Bull., vol. 68, pp. 203–209, 2015. https://doi.org/10.1016/j.materresbull.2015.03.064.Search in Google Scholar
[95] D. Kumar, C. H. Park, and C. S. Kim, “Robust multimetallic plasmonic core–satellite nanodendrites: highly effective visible-light-induced colloidal CO2 photoconversion system,” ACS Sustainable Chem. Eng., vol. 6, no. 7, pp. 8604–8614, 2018. https://doi.org/10.1021/acssuschemeng.8b00924.Search in Google Scholar
[96] D. Kumar, A. Lee, T. Lee, M. Lim, and D. K. Lim, “Ultrafast and efficient transport of hot plasmonic electrons by graphene for Pt free, highly efficient visible-light responsive photocatalyst,” Nano Lett., vol. 16, no. 3, pp. 1760–1767, 2016. https://doi.org/10.1021/acs.nanolett.5b04764.Search in Google Scholar PubMed
[97] J. Zhao, B. Liu, L. Meng, et al.., “Plasmonic control of solar-driven CO2 conversion at the metal/ZnO interfaces,” Appl. Catal., B, vol. 256, p. 117823, 2019. https://doi.org/10.1016/j.apcatb.2019.117823.Search in Google Scholar
[98] X. Zhang, X. Li, M. E. Reish, et al.., “Plasmon-enhanced catalysis: distinguishing thermal and nonthermal effects,” Nano Lett., vol. 18, no. 3, pp. 1714–1723, 2018. https://doi.org/10.1021/acs.nanolett.7b04776.Search in Google Scholar PubMed
[99] H. Robatjazi, D. Weinberg, D. F. Swearer, et al.., “Metal-organic frameworks tailor the properties of aluminum nanocrystals,” Sci. Adv., vol. 5, no. 2, p. eaav5340, 2019. https://doi.org/10.1126/sciadv.aav5340.Search in Google Scholar PubMed PubMed Central
[100] E. Prodan, C. Radloff, N. J. Halas, and P. Nordlander, “A hybridization model for the plasmon response of complex nanostructures,” Science, vol. 302, no. 5644, pp. 419–422, 2003. https://doi.org/10.1126/science.1089171.Search in Google Scholar PubMed
[101] B. Y. Zheng, H. Zhao, A. Manjavacas, M. McClain, P. Nordlander, and N. J. Halas, “Distinguishing between plasmon-induced and photoexcited carriers in a device geometry,” Nat. Commun., vol. 6, p. 7797, 2015. https://doi.org/10.1038/ncomms8797.Search in Google Scholar PubMed PubMed Central
[102] H. Robatjazi, M. Lou, B. D. Clark, et al.., “Site-selective nanoreactor deposition on photocatalytic Al nanocubes,” Nano Lett., vol. 20, no. 6, pp. 4550–4557, 2020. https://doi.org/10.1021/acs.nanolett.0c01405.Search in Google Scholar PubMed
[103] H. Robatjazi, J. L. Bao, M. Zhang, et al.., “Plasmon-driven carbon–fluorine (C(sp3)–F) bond activation with mechanistic insights into hot-carrier-mediated pathways,” Nat. Catal., vol. 3, no. 7, pp. 564–573, 2020. https://doi.org/10.1038/s41929-020-0466-5.Search in Google Scholar
[104] A. Al-Zubeidi, B. Ostovar, C. C. Carlin, et al.., “Mechanism for plasmon-generated solvated electrons,” Proc. Natl. Acad. Sci. U. S. A., vol. 120, no. 3, 2023, Art. no. e2217035120. https://doi.org/10.1073/pnas.2217035120.Search in Google Scholar PubMed PubMed Central
[105] L. V. Besteiro, X.-T. Kong, Z. Wang, G. Hartland, and A. O. Govorov, “Understanding hot-electron generation and plasmon relaxation in metal nanocrystals: quantum and classical mechanisms,” ACS Photonics, vol. 4, no. 11, pp. 2759–2781, 2017. https://doi.org/10.1021/acsphotonics.7b00751.Search in Google Scholar
[106] A. D. Rakic, A. B. Djurisic, J. M. Elazar, and M. L. Majewski, “Optical properties of metallic films for vertical-cavity optoelectronic devices,” Appl. Opt., vol. 37, no. 22, pp. 5271–5283, 1998. https://doi.org/10.1364/ao.37.005271.Search in Google Scholar PubMed
[107] J. M. Rahm, C. Tiburski, T. P. Rossi, et al.., “A library of late transition metal alloy dielectric functions for nanophotonic applications,” Adv. Funct. Mater., vol. 30, no. 35, p. 2002122, 2020. https://doi.org/10.1002/adfm.202002122.Search in Google Scholar
[108] L. Yuan, J. Zhou, M. Zhang, et al.., “Plasmonic photocatalysis with chemically and spatially specific antenna-dual reactor complexes,” ACS Nano, vol. 16, no. 10, pp. 17365–17375, 2022. https://doi.org/10.1021/acsnano.2c08191.Search in Google Scholar PubMed
[109] C. Frischkorn and M. Wolf, “Femtochemistry at metal surfaces: nonadiabatic reaction dynamics,” Chem. Rev., vol. 106, no. 10, pp. 4207–4233, 2006. https://doi.org/10.1021/cr050161r.Search in Google Scholar PubMed
[110] A. Varas, P. García-González, J. Feist, F. J. García-Vidal, and A. Rubio, “Quantum plasmonics: from jellium models to ab initio calculations,” Nanophotonics, vol. 5, no. 3, pp. 409–426, 2016. https://doi.org/10.1515/nanoph-2015-0141.Search in Google Scholar
[111] M. Kuisma, A. Sakko, T. P. Rossi, et al.., “Localized surface plasmon resonance in silver nanoparticles: atomistic first-principles time-dependent density-functional theory calculations,” Phys. Rev. B, vol. 91, no. 11, p. 115431, 2015. https://doi.org/10.1103/PhysRevB.91.115431.Search in Google Scholar
[112] T. P. Rossi, M. Kuisma, M. J. Puska, R. M. Nieminen, and P. Erhart, “Kohn–Sham decomposition in real-time time-dependent density-functional theory: an efficient tool for analyzing plasmonic excitations,” J. Chem. Theory Comput., vol. 13, no. 10, pp. 4779–4790, 2017. https://doi.org/10.1021/acs.jctc.7b00589.Search in Google Scholar PubMed
[113] P. V. Kumar, T. P. Rossi, D. Marti-Dafcik, et al.., “Plasmon-induced direct hot-carrier transfer at metal–acceptor interfaces,” ACS Nano, vol. 13, no. 3, pp. 3188–3195, 2019. https://doi.org/10.1021/acsnano.8b08703.Search in Google Scholar PubMed
[114] K. M. Conley, N. Nayyar, T. P. Rossi, et al.., “Plasmon excitations in mixed metallic nanoarrays,” ACS Nano, vol. 13, no. 5, pp. 5344–5355, 2019. https://doi.org/10.1021/acsnano.8b09826.Search in Google Scholar PubMed
[115] P. V. Kumar, T. P. Rossi, M. Kuisma, P. Erhart, and D. J. Norris, “Direct hot-carrier transfer in plasmonic catalysis,” Faraday Discuss., vol. 214, pp. 189–197, 2019. https://doi.org/10.1039/C8FD00154E.Search in Google Scholar PubMed
[116] J. Ma, Z. Wang, and L.-W. Wang, “Interplay between plasmon and single-particle excitations in a metal nanocluster,” Nat. Comm., vol. 6, no. 1, p. 10107, 2015. https://doi.org/10.1038/ncomms10107.Search in Google Scholar PubMed PubMed Central
[117] L. R. Castellanos, J. M. Kahk, O. Hess, and J. Lischner, “Generation of plasmonic hot carriers from d-bands in metallic nanoparticles,” J. Chem. Phys., vol. 152, no. 10, p. 104111, 2020. https://doi.org/10.1063/5.0003123.Search in Google Scholar PubMed
[118] T. P. Rossi, P. Erhart, and M. Kuisma, “Hot-carrier generation in plasmonic nanoparticles: the importance of atomic structure,” ACS Nano, vol. 14, no. 8, pp. 9963–9971, 2020. https://doi.org/10.1021/acsnano.0c03004.Search in Google Scholar PubMed PubMed Central
[119] R. Sundararaman, P. Narang, A. S. Jermyn, W. A. GoddardIii, and H. A. Atwater, “Theoretical predictions for hot-carrier generation from surface plasmon decay,” Nat. Comm., vol. 5, no. 1, p. 5788, 2014. https://doi.org/10.1038/ncomms6788.Search in Google Scholar PubMed PubMed Central
[120] M. Bernardi, J. Mustafa, J. B. Neaton, and S. G. Louie, “Theory and computation of hot carriers generated by surface plasmon polaritons in noble metals,” Nat. Comm., vol. 6, no. 1, p. 7044, 2015. https://doi.org/10.1038/ncomms8044.Search in Google Scholar PubMed PubMed Central
[121] S. Mukherjee, F. Libisch, N. Large, et al.., “Hot electrons do the impossible: plasmon-induced dissociation of H2 on Au,” Nano Lett., vol. 13, no. 1, pp. 240–247, 2013. https://doi.org/10.1021/nl303940z.Search in Google Scholar PubMed
[122] C. Schäfer, J. Flick, E. Ronca, P. Narang, and A. Rubio, “Shining light on the microscopic resonant mechanism responsible for cavity-mediated chemical reactivity,” Nat. Comm., vol. 13, no. 1, p. 7817, 2022. https://doi.org/10.1038/s41467-022-35363-6.Search in Google Scholar PubMed PubMed Central
[123] M. Del Ben, F. H. da Jornada, A. Canning, et al.., “Large-scale GW calculations on pre-exascale HPC systems,” Comput. Phys. Commun., vol. 235, pp. 187–195, 2019. https://doi.org/10.1016/j.cpc.2018.09.003.Search in Google Scholar
[124] M. D. Ben, C. Yang, Z. Li, F. H. D. Jornada, S. G. Louie, and J. Deslippe, “Accelerating large-scale excited-state GW Calculations on leadership HPC systems,” in SC20: International Conference for High Performance Computing, Networking, Storage and Analysis, vol. 2020, pp. 1–11, 2020.10.1109/SC41405.2020.00008Search in Google Scholar
[125] V. W.-Z. Yu and M. Govoni, “GPU acceleration of large-scale full-frequency GW calculations,” J. Chem. Theory Comput., vol. 18, no. 8, pp. 4690–4707, 2022. https://doi.org/10.1021/acs.jctc.2c00241.Search in Google Scholar PubMed
[126] J. Wilhelm, D. Golze, L. Talirz, J. Hutter, and C. A. Pignedoli, “Toward GW calculations on thousands of atoms,” J. Phys. Chem. Lett., vol. 9, no. 2, pp. 306–312, 2018. https://doi.org/10.1021/acs.jpclett.7b02740.Search in Google Scholar PubMed
[127] M. S. Hybertsen and S. G. Louie, “Electron correlation in semiconductors and insulators: band gaps and quasiparticle energies,” Phys. Rev. B, vol. 34, no. 8, pp. 5390–5413, 1986. https://doi.org/10.1103/PhysRevB.34.5390.Search in Google Scholar
[128] M. Rohlfing and S. G. Louie, “Electron-hole excitations and optical spectra from first principles,” Phys. Rev. B, vol. 62, no. 8, pp. 4927–4944, 2000. https://doi.org/10.1103/PhysRevB.62.4927.Search in Google Scholar
[129] D. K. Angell, B. Bourgeois, M. Vadai, and J. A. Dionne, “Lattice-resolution, dynamic imaging of hydrogen absorption into bimetallic AgPd nanoparticles,” ACS Nano, vol. 16, no. 2, pp. 1781–1790, 2022. https://doi.org/10.1021/acsnano.1c04602.Search in Google Scholar PubMed
[130] M. Vadai, D. K. Angell, F. Hayee, K. Sytwu, and J. A. Dionne, “In-situ observation of plasmon-controlled photocatalytic dehydrogenation of individual palladium nanoparticles,” Nat. Commun., vol. 9, no. 1, p. 4658, 2018. https://doi.org/10.1038/s41467-018-07108-x.Search in Google Scholar PubMed PubMed Central
[131] K. Sytwu, M. Vadai, F. Hayee, et al.., “Driving energetically unfavorable dehydrogenation dynamics with plasmonics,” Science, vol. 371, no. 6526, pp. 280–283, 2021. https://doi.org/10.1126/science.abd2847.Search in Google Scholar PubMed
© 2023 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Frontmatter
- Editorial
- Nanophotonics in support of Ukrainian Scientists
- Reviews
- Asymmetric transmission in nanophotonics
- Integrated circuits based on broadband pixel-array metasurfaces for generating data-carrying optical and THz orbital angular momentum beams
- Singular optics empowered by engineered optical materials
- Electrochemical photonics: a pathway towards electrovariable optical metamaterials
- Sustainable chemistry with plasmonic photocatalysts
- Perspectives
- Ukraine and singular optics
- Machine learning to optimize additive manufacturing for visible photonics
- Through thick and thin: how optical cavities control spin
- Research Articles
- Spin–orbit coupling induced by ascorbic acid crystals
- Broadband transfer of binary images via optically long wire media
- Counting and mapping of subwavelength nanoparticles from a single shot scattering pattern
- Controlling surface waves with temporal discontinuities of metasurfaces
- On the relation between electrical and electro-optical properties of tunnelling injection quantum dot lasers
- On-chip multivariant COVID 19 photonic sensor based on silicon nitride double-microring resonators
- Nano-infrared imaging of metal insulator transition in few-layer 1T-TaS2
- Electrical generation of surface phonon polaritons
- Dynamic beam control based on electrically switchable nanogratings from conducting polymers
- Tilting light’s polarization plane to spatially separate the ultrafast nonlinear response of chiral molecules
- Spin-dependent phenomena at chiral temporal interfaces
- Spin-controlled photonics via temporal anisotropy
- Coherent control of symmetry breaking in transverse-field Ising chains using few-cycle pulses
- Field enhancement of epsilon-near-zero modes in realistic ultrathin absorbing films
- Controlled compression, amplification and frequency up-conversion of optical pulses by media with time-dependent refractive index
- Tailored thermal emission in bulk calcite through optic axis reorientation
- Tip-enhanced photoluminescence of monolayer MoS2 increased and spectrally shifted by injection of electrons
- Quantum-enhanced interferometer using Kerr squeezing
- Nonlocal electro-optic metasurfaces for free-space light modulation
- Dispersion braiding and band knots in plasmonic arrays with broken symmetries
- Dual-mode hyperbolicity, supercanalization, and leakage in self-complementary metasurfaces
- Monocular depth sensing using metalens
- Multimode hybrid gold-silicon nanoantennas for tailored nanoscale optical confinement
- Replicating physical motion with Minkowskian isorefractive spacetime crystals
- Reconfigurable nonlinear optical element using tunable couplers and inverse-designed structure
Articles in the same Issue
- Frontmatter
- Editorial
- Nanophotonics in support of Ukrainian Scientists
- Reviews
- Asymmetric transmission in nanophotonics
- Integrated circuits based on broadband pixel-array metasurfaces for generating data-carrying optical and THz orbital angular momentum beams
- Singular optics empowered by engineered optical materials
- Electrochemical photonics: a pathway towards electrovariable optical metamaterials
- Sustainable chemistry with plasmonic photocatalysts
- Perspectives
- Ukraine and singular optics
- Machine learning to optimize additive manufacturing for visible photonics
- Through thick and thin: how optical cavities control spin
- Research Articles
- Spin–orbit coupling induced by ascorbic acid crystals
- Broadband transfer of binary images via optically long wire media
- Counting and mapping of subwavelength nanoparticles from a single shot scattering pattern
- Controlling surface waves with temporal discontinuities of metasurfaces
- On the relation between electrical and electro-optical properties of tunnelling injection quantum dot lasers
- On-chip multivariant COVID 19 photonic sensor based on silicon nitride double-microring resonators
- Nano-infrared imaging of metal insulator transition in few-layer 1T-TaS2
- Electrical generation of surface phonon polaritons
- Dynamic beam control based on electrically switchable nanogratings from conducting polymers
- Tilting light’s polarization plane to spatially separate the ultrafast nonlinear response of chiral molecules
- Spin-dependent phenomena at chiral temporal interfaces
- Spin-controlled photonics via temporal anisotropy
- Coherent control of symmetry breaking in transverse-field Ising chains using few-cycle pulses
- Field enhancement of epsilon-near-zero modes in realistic ultrathin absorbing films
- Controlled compression, amplification and frequency up-conversion of optical pulses by media with time-dependent refractive index
- Tailored thermal emission in bulk calcite through optic axis reorientation
- Tip-enhanced photoluminescence of monolayer MoS2 increased and spectrally shifted by injection of electrons
- Quantum-enhanced interferometer using Kerr squeezing
- Nonlocal electro-optic metasurfaces for free-space light modulation
- Dispersion braiding and band knots in plasmonic arrays with broken symmetries
- Dual-mode hyperbolicity, supercanalization, and leakage in self-complementary metasurfaces
- Monocular depth sensing using metalens
- Multimode hybrid gold-silicon nanoantennas for tailored nanoscale optical confinement
- Replicating physical motion with Minkowskian isorefractive spacetime crystals
- Reconfigurable nonlinear optical element using tunable couplers and inverse-designed structure