Abstract
Artificial photosynthesis of hydrocarbons from carbon dioxide (CO2) has the potential to provide renewable fuels at the scale needed to meet global decarbonization targets. However, CO2 is a notoriously inert molecule and converting it to energy dense hydrocarbons is a complex, multistep process, which can proceed through several intermediates. Recently, the ability of plasmonic nanoparticles to steer the reaction down specific pathways and enhance both reaction rate and selectivity has garnered significant attention due to its potential for sustainable energy production and environmental mitigation. The plasmonic excitation of strong and confined optical near-fields, energetic hot carriers and localized heating can be harnessed to control or enhance chemical reaction pathways. However, despite many seminal contributions, the anticipated transformative impact of plasmonics in selective CO2 photocatalysis has yet to materialize in practical applications. This is due to the lack of a complete theoretical framework on the plasmonic action mechanisms, as well as the challenge of finding efficient materials with high scalability potential. In this review, we aim to provide a comprehensive and critical discussion on recent advancements in plasmon-enhanced CO2 photoreduction, highlighting emerging trends and challenges in this field. We delve into the fundamental principles of plasmonics, discussing the seminal works that led to ongoing debates on the reaction mechanism, and we introduce the most recent ab initio advances, which could help disentangle these effects. We then synthesize experimental advances and in situ measurements on plasmon CO2 photoreduction before concluding with our perspective and outlook on the field of plasmon-enhanced photocatalysis.
1 Background and motivation
The ever-accelerating pace of industrialization and anthropogenic activities has led to an alarming increase in atmospheric CO2 levels over the past 100 years, reaching >420 ppm in 2023 [1], fuelling concerns about climate change and its far-reaching consequences. Addressing this critical environmental challenge demands innovative and multifaceted approaches that not only reduce CO2 emissions but also transform them into valuable chemicals and fuels. Particularly, hydrocarbons with more than one carbon atom per molecule (C2+) are valuable due to their high energy density, versatility, ease of storage and transportation and large, established market. Unfortunately, producing these chemicals using current methods relying on fossil fuels emits millions of tonnes of CO2-equivalent greenhouse gases per year.
Artificial photosynthesis enables a closed-loop solution for producing hydrocarbons from CO2 and water using sunlight to drive the reaction. If the CO2 is directly captured from the atmosphere, this has the potential to provide renewable fuels at scale with net-zero emissions, without the need for additional electricity inputs. Upon light absorption, photocatalytic materials including semiconductors and metals generate electron–hole pairs, which migrate to the catalyst’s surface and can be used to drive two related redox half reactions: electrons for reducing CO2 and holes for oxidizing water [2], [3]. While both sides of the reactions are important for a complete treatment of artificial photosynthesis, the CO2 reduction is a more complex and harder to realize.
Due to its stability, CO2 reduction into higher-value products is a complex, multistep process requiring multiple electron and proton transfers. As shown in Scheme 1, the existence of multiple, branching and co-existing reaction pathways can lead to a variety of different hydrocarbons. The challenge in achieving commercially viable renewable production of valuable (C2+) hydrocarbon products using solar energy and CO2 is selectively producing the desired product with high yields. Ultimately, this depends on the physicochemical properties of the materials and is a critical object of research. For this reason, a large body of recent work has focused exclusively on the rational design of efficient photocatalysts for CO2 reduction half reaction, and studies often use hole scavengers as sacrificial electron donors, to prevent surface hole accumulation and improve charge separation [2], [4], [5], [6].
![Scheme 1:
Mechanistic pathways of CO2 reduction to commonly observed C1 and C2 species on metal surfaces. Reproduced with permission from ref [7]. Copyright 2022, American Chemical Society.](/document/doi/10.1515/nanoph-2023-0793/asset/graphic/j_nanoph-2023-0793_scheme_001.jpg)
Mechanistic pathways of CO2 reduction to commonly observed C1 and C2 species on metal surfaces. Reproduced with permission from ref [7]. Copyright 2022, American Chemical Society.
Leveraging the exquisite control of light–matter interactions at the nanoscale afforded by plasmonic nanostructures offers a prospective avenue for efficiently facilitating chemical transformations. Recently, the use of plasmonic materials as active elements has sparked renewed interest the field of plasmonics, thanks to the demonstration of enhanced photochemical reaction and the promise of achieving reaction selectivity [4], [8], [9], [10], [11], [12], [13], suggesting a strategy for sustainable chemical production.
Upon light interaction, nanostructured metals exhibit strong light–matter interactions, resulting in the excitation of localized surface plasmon resonances (LSPRs). These provide a way to concentrate light in subwavelength volumes resulting in highly enhanced electric fields strongly confined at the nanostructure’s surface. Following their excitation, the non-radiative decay of the LSPRs generates a population of highly energetic electron–hole pairs, commonly referred to as hot carriers. These photo-excited carriers gradually lose their energy through scattering events and ultimately release it to the environment as heat, resulting in a localized temperature increase. Enhanced fields, hot carriers and thermal gradient – either individually or synergistically – can be harnessed to drive and promote chemical reactions [14], [15], [16], [17], [18], [19]. Nonetheless, despite seminal contributions and proof-of-concept demonstrations, the anticipated transformative impact of plasmonics on the field of photocatalysis has yet to materialize in practical applications.
In this review, we will elucidate the challenges and address the opportunities in realizing practical applications of plasmon-mediated CO2 photocatalysis. We aim to provide a comprehensive overview of the recent advancements in plasmonic photocatalysis for CO2 reduction, focusing on the key mechanisms, challenges and prospects. For a thorough exploration of the other half-reaction and an in-depth discussion on the role of holes in chemical reactions, we refer the readers to excellent recent reviews on the topic [20], [21], [22]. We will start by discussing the fundamentals of plasmon photocatalysis and the seminal works that have led to the ongoing debate on the reaction mechanism. In doing so, we will clarify the multiple adopted definitions and nomenclature. We will then discuss the most recent theoretical advances in understanding the opto-electronic properties of plasmonic materials and how they interface with the catalysis field, paying particular attention to their coupling and interactions with CO2. Here, we will introduce ab initio density functional theory calculations, along with their challenges and perspectives. Next, we will discuss recent experimental demonstrations of plasmonic CO2 photoreduction, differentiating between in situ measurement, often used for investigation of the plasmon action mechanisms, and ensemble measurements, aimed at demonstrating high reaction yields, efficiency and selectivity. Here, we will analyse how the theoretical results discussed in the previous section support and/or contrast the experimental findings. Finally, we will provide our point of view on the development of this field for scalable renewable fuel production from CO2 splitting. We will conclude by discussing the challenges and opportunities to realize the practical application of plasmonic photocatalysis for large scale renewable fuel production.
1.1 Plasmon excitation and relaxation
In the past few decades, the rapid development of nanotechnology has enabled subwavelength structuring of materials, allowing the control of light–matter interactions at optical wavelengths [23], [24], [25], [26], [27]. Particular attention has been given to metallic nanoparticles, due to their ability to strongly interact and couple with photons of suitable energy, their ease of fabrication and the ability to easily model their optical behaviour [28], [29], [30].
Within illuminated metal nanostructures, the alternating electric component of incident electromagnetic waves induces a coherent displacement of the metal free electron density, which is counterbalanced by a restoring coulombic force between the nuclei and the electrons. This results in the collective oscillations of the metal conduction electrons, known as localized surface plasmon resonance (LSPR) [31]. When resonantly excited, metallic nanoparticles (NPs) can interact with incident light over areas much larger than their own geometric cross section and can absorb and concentrate light in subwavelength volumes [32], effectively acting as nanoscale optical antennas [31], [33]. This interaction results in light energy confinement in the form of enhanced electromagnetic fields at the surface of the metal nanostructure [34], [35], which are localized within the first ∼30 nm from the nanoparticle’s surface (Figure 1(a)) [36]. These are commonly referred to as near-fields and are heavily exploited in surface-enhanced spectroscopies [32], [37], [38].
![Figure 1:
Plasmon relaxation process and energy transfer mechanisms. (a) Main effects resulting from plasmon excitation and relaxation schematically represented in the space, energy and time domains. These are classified in non-thermal (electromagnetic near-fields and hot carriers) and thermal effects (local heating). (b-i) Schematic of the possible plasmon energy transfer mechanisms in the space domain. (b-ii) Plasmon action mechanisms as a function of the energy and reaction coordinate. R and P are the reactant and product molecules, respectively, while φ
G and φ
E represents the ground and an excited state PESs, respectively. Near-fields and hot carriers can promote the system to an excited PES, while local heat can provide only vibrational energy along φ
G. (b-iii) Hybridization between molecular and metallic states can decrease the energy gap required for transferring hot electrons. These can be transferred directly (blue arrow) or indirectly (red arrows). (c) Rendering of plasmon-enabled selectivity via promotion of the system from the ground to an excited potential energy surface. The excited state may have lower activation barrier for product generation (higher efficiency) and different minima (linked to selectivity). (a) Reproduced with permission from [14]. Copyright 2020, Elsevier. (b-i, b-iii) Adapted with permission from [17]. Copyright 2022, The Royal Society of Chemistry. (b-ii) Reproduced with permission from [14]. Reproduced with permission from [14]. Copyright 2020, Elsevier. (c) Reproduced with permission from [18]. Copyright 2019, American Chemical Society.](/document/doi/10.1515/nanoph-2023-0793/asset/graphic/j_nanoph-2023-0793_fig_001.jpg)
Plasmon relaxation process and energy transfer mechanisms. (a) Main effects resulting from plasmon excitation and relaxation schematically represented in the space, energy and time domains. These are classified in non-thermal (electromagnetic near-fields and hot carriers) and thermal effects (local heating). (b-i) Schematic of the possible plasmon energy transfer mechanisms in the space domain. (b-ii) Plasmon action mechanisms as a function of the energy and reaction coordinate. R and P are the reactant and product molecules, respectively, while φ G and φ E represents the ground and an excited state PESs, respectively. Near-fields and hot carriers can promote the system to an excited PES, while local heat can provide only vibrational energy along φ G. (b-iii) Hybridization between molecular and metallic states can decrease the energy gap required for transferring hot electrons. These can be transferred directly (blue arrow) or indirectly (red arrows). (c) Rendering of plasmon-enabled selectivity via promotion of the system from the ground to an excited potential energy surface. The excited state may have lower activation barrier for product generation (higher efficiency) and different minima (linked to selectivity). (a) Reproduced with permission from [14]. Copyright 2020, Elsevier. (b-i, b-iii) Adapted with permission from [17]. Copyright 2022, The Royal Society of Chemistry. (b-ii) Reproduced with permission from [14]. Reproduced with permission from [14]. Copyright 2020, Elsevier. (c) Reproduced with permission from [18]. Copyright 2019, American Chemical Society.
Once excited, plasmons can decay radiatively by photon emission or non-radiatively by generating a population of highly energetic charge carriers. The latter are out-of-equilibrium electron–hole pairs generated within 100 fs with an energy distribution that cannot be described by the Fermi–Dirac statistics and are commonly referred to as hot carriers [39], [40]. Recent theoretical investigation on the spatial distribution of hot carriers within small Ag clusters revealed that hot holes accumulated at atomic sites throughout the particle, while hot electrons reside mainly in the surface region [41]. Utilization of these hot carriers often requires charge separation, which can be achieved for instance by forming a junction at metal–semiconductor interfaces [42], [43], [44], [45]. Within hundreds of fs, the hot carriers are subjected to internal electron–electron scattering events, which results in a gradual redistribution of their energy and establishment of a population of carriers in thermal equilibrium with a well-defined electronic temperature (T e), higher than that of the lattice, T L. This process is commonly referred to as thermalization and can be described by using the so-called two-temperature model [46], [47]; at this stage, the energy distribution of the hot carriers can be described by a Fermi–Dirac distribution. Further electron–phonon scattering leads to equilibration between T e and T L within few picoseconds and results in an increased temperature of the metal [46], [48]. While the photophysical process for hot electron and hot holes generation is the same, it is worth noting that the photo-generated hot holes usually have a shorter lifetime than hot-electrons [41], [49], [50], making them harder to employ in driving redox reactions [20]. Ultimately, within nanoseconds, the heat is transferred to the environment through phonon–phonon interactions, resulting in a macroscopic temperature increase [14], [48]. Figure 1(b) schematically shows these relaxation effects.
1.2 Plasmon energy transfer mechanisms
During the excitation and relaxation processes of the LSPRs, three different mechanisms transfer energy from the optical source to surrounding molecular species or semiconductors via the plasmonic nanostructure: enhanced near-fields, hot carriers and heat, as shown in Figure 1(b, i-ii). These transfer processes can be classified into non-thermal and thermal effects, where the former include near-field enhancements and hot carrier generation, while the latter refers to the local heating effects [17], [19], [51], [52]. Some authors have also referred these effects as photochemical and photothermal, respectively [11], [53], [54].
1.2.1 Near-field enhancement
Near-field enhancements can induce optical transitions of adjacent molecules via excitation between highest unoccupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) [11], [33], [55], [56], [57], promoting the catalyst–reactant system, (consisting of the metal nanostructured catalysts and an absorbed molecular species), to an excited potential energy surface (PES), as shown in by ψ E in Figure 1(b-ii). This only takes place when there is spectral overlap between the LSPR and the molecular energy gap of the acceptor. Increase in the intensity of these fields results in higher probability of optical transitions, thus engineering the optical properties of metallic nanostructures to achieve regions with high field intensity – referred to as hot spots – is desirable to increase the reaction rate [58]. In addition, as the near-fields are strongest at the NPs surface, reactant molecules should be within the evanescent volume of the plasmon fields for maximum efficiency [59], [60]. The rate of near-field induced transitions can be enhanced by the re-emitted photons from the radiative plasmon decay process [33]. Given the high multi-disciplinarity of the field, this process has also been referred to as plasmon-induced intramolecular excitation, direct intramolecular excitation, plasmon-pumped adsorbate excitation, adsorbate electronic excitation or plasmon-induced resonance energy transfer [59], [61].
1.2.2 Hot carriers
The second non-thermal mechanism involves hot carriers. These are preferentially generated in regions where the electric near-field is strongest [62], [63], and during their short lifetime, they can be transiently transferred to accepting orbitals of nearby molecules. Provided that they have suitable energy and alignment with adsorbate orbitals, hot electrons (holes) can be transferred to the LUMO (HOMO) of adsorbed molecular species, driving reduction (oxidation) reactions. Generally, this transfer is reversible, resulting in the creation of a transient negative ion (TNI), which promotes the system along an electronically excited potential energy surface (ψ E in Figure 1(b-ii)), thereby catalysing chemical reactions [10], [64], [65], [66], [67], [68]. Extending the lifetime of hot carriers within the acceptor orbitals of reactant molecules is a promising strategy to increase the probability of inducing a chemical reaction. Hot electron charge transfer can also initiate bond formation or drive redox reactions, in which case the ion or radical disassociates and the electron transfer is irreversible. Transfer can occur indirectly (Figure 1(b-iii), red arrows), where carriers are generated within the metal nanoparticles and then scattered into accessible molecular orbitals, provided they satisfy energy and momentum requirements [55], [69]. However, it is also possible for the energy levels of molecule–metal adsorbates to hybridize, allowing the direct excitation (Figure 1(b-iii), blue arrow) of hot carriers into these newly hybridized surface states [55], [69]. Hybridization affects the energy of the HOMO–LUMO, often leading to a reduced molecular gap (Figure 1(c-iii)) and opens a new energy transfer channel for hot carrier relaxation. Direct transfer, also referred to as chemical interface damping (CID) or plasmon resonant energy transfer [70], is a faster and more efficient process compared to the indirect transfer [33], [71], as it avoids energy losses due to scattering within the metal and during the injection process across the interface and is especially promising in terms of increased performance [33], [72], [73].
One of the promises of plasmonic photocatalysis is the possibility of achieving product selectivity by activating reaction pathways that are not attainable with conventional thermocatalysis [11], [73]. This can be achieved by providing energy via hot carrier transfer or near-fields to specific unoccupied adsorbate states, thereby promoting the system along an excited PES and facilitating targeted reaction pathways [15], [18], [73]. As shown in Figure 1(c), accessing excited states results in the modification of the reaction pathway corresponding to the production of different products. As near-fields and hot electrons energies depend on the physical properties of the plasmon nanostructures, the design and engineering of their optoelectronic properties is particularly important to achieve selectivity in multiproduct reactions such as CO2 reduction [74].
1.2.3 Photothermal effect
Ultimately, hot carriers that not transferred to adsorbates will gradually lose their energy via scattering events releasing their energy as heat, thus leading to an increase in local temperature. This can enhance chemical transformation by increasing the population of reactants in vibrationally excited states (red arrows in Figure 1(c-ii)) and hence accelerating the reaction rate. As the reaction rate benefits from an increase in temperature, this energy transfer mechanism can be advantageous in many reactions for achieving higher product yields; however, as the photothermal effect drives the catalytic reaction along the ground state PES, it does not enable selectivity.
The temperature increase in a plasmonic system depends on the optoelectronic properties of the nanoparticles, their size, concentration and distribution of as well as the thermal properties of the surrounding medium and illumination conditions [1]. Taking into account these variables, the resultant temperature increase can be either confined in the vicinity of the plasmonic structures or – if collective effects are present – delocalized. In the latter case, the temperature profile is uniform throughout the whole nanoparticle assembly.
Given that the very fast timescales involved in the plasmon excitation and relaxation processes are similar to reaction time constants, disentangling thermal from non-thermal effects and understanding their relative contribution to plasmon reactivity has been object of a fruitful debate, as discussed in the next section [70], [75], [76], [77], [78].
1.3 Ongoing debate: photochemical versus photothermal
The primary subject of controversy is the elusive differentiation between thermal and non-thermal effects [19], [52], [54], [79]. One way to differentiate them would be to monitor local temperature increases, calculate the expected photothermal effect and deduce its impact on reaction rates using the well-known Arrhenius law. However, although it can seem a straightforward task, measuring plasmon-induced temperature increases on the surface of illuminated nanostructures is a challenging exercise because of the fast kinetics and small length scales involved, the complex light absorption and scattering pathways that affects mass and heat transport and large temperature gradients. When combined with factors relating to instrumental sensitivity, spatial resolution and uncertainty, it becomes apparent why this challenge proves demanding and has contributed to the ongoing debate.
The intricate interplay between thermal and non-thermal mechanisms has resulted in contrasting outcomes and divergent interpretations of data [80], [81]. Zhou et al. conducted a seminal study to quantify the contribution of hot carriers and thermal effects in enhancing ammonia (NH3) decomposition using Cu–Ru photocatalysts [64]. The authors measured the reaction rate under light illumination and in dark conditions with external heating and used a thermocouple to monitor the temperature of the sample surface. The results revealed a considerably higher reaction yield for the illuminated sample at comparable measured temperatures to those in the dark. By fitting the reaction rate with an Arrhenius law, the authors demonstrated that the activation energy is intensity and wavelength dependent under illumination (E a (I inc, λ )) and reduces from 1.21 eV in the dark, to 0.35 eV at resonance, as shown in Figure 2(a). They concluded that thermal effects are not sufficient to explain the observed behaviour, and that the reaction is catalysed by plasmon-induced hot carriers. Nevertheless, the temperature evaluation within this approach was criticized: Dubi and colleagues suggested that the temperature measured by the thermocouple may not accurately capture the reaction conditions on the surface of the plasmonic photocatalysis. They went on to show that the reaction rates seen experimentally could be modelled using a purely thermal approach by fitting an Arrhenius equation with a constant activation energy (E a) and instead assuming an intensity dependent increase in the reaction temperature under illumination (∆T = aI inc, λ ) (Figure 2(b)) [80]. This highlights that without an accurate measurement of the local reaction temperature, fitting experimental data to an Arrhenius equation does not provide conclusive proof that a chemical reaction is driven either by photochemical or photothermal effects [82].
![Figure 2:
Photochemical versus photothermal debate. (a) Light-dependent activation energy, E
a(I
inc, λ
) for the NH3 decomposition on Cu–Ru photocatalysts. (b) Reaction rate enhancement factor under plasmonic excitation fitted with a light-dependent activation energy and constant temperature (photochemical model, red dots) and with a light-dependent temperature but constant activation energy (photothermal model, black line). (c) Yield of CO2 transformed to methane on Rh–TiO2 photocatalysts, using direct (orange triangles) and indirect (red circles) illumination. The purple line represents the calculated thermal yield, while the shaded orange area represents the non-thermal contributions. (d) Enhancement factors (calculated by dividing the photothermal rates by respective thermal rates) as a function of the external heat supply for the CO oxidation on Ag/Al2O3 catalysts with different mass loadings. (a) Reproduced with permission from [64]. Copyright 2018, The American Association for the Advancement of Science. (b) Reproduced with permission from ref. [82]. Copyright 2020, Royal Society of Chemistry. (c) Reproduced with permission from [83]. Copyright 2019, Springer Nature. (d) Reproduced with permission from [84]. Copyright 2022, American Chemical Society.](/document/doi/10.1515/nanoph-2023-0793/asset/graphic/j_nanoph-2023-0793_fig_002.jpg)
Photochemical versus photothermal debate. (a) Light-dependent activation energy, E a(I inc, λ ) for the NH3 decomposition on Cu–Ru photocatalysts. (b) Reaction rate enhancement factor under plasmonic excitation fitted with a light-dependent activation energy and constant temperature (photochemical model, red dots) and with a light-dependent temperature but constant activation energy (photothermal model, black line). (c) Yield of CO2 transformed to methane on Rh–TiO2 photocatalysts, using direct (orange triangles) and indirect (red circles) illumination. The purple line represents the calculated thermal yield, while the shaded orange area represents the non-thermal contributions. (d) Enhancement factors (calculated by dividing the photothermal rates by respective thermal rates) as a function of the external heat supply for the CO oxidation on Ag/Al2O3 catalysts with different mass loadings. (a) Reproduced with permission from [64]. Copyright 2018, The American Association for the Advancement of Science. (b) Reproduced with permission from ref. [82]. Copyright 2020, Royal Society of Chemistry. (c) Reproduced with permission from [83]. Copyright 2019, Springer Nature. (d) Reproduced with permission from [84]. Copyright 2022, American Chemical Society.
Similar discrepancies [80], [85] are found in other works that have tried to discern between thermal and non-thermal effects for different reactions by means of standard macroscopic methodologies and surface-enhanced spectroscopies [86], [87], [88]. For instance, the plasmon-driven reduction of pNTP has been claimed to be driven both from non-thermal [89] and thermal effects [90]. Similarly, the N-demethylation reaction of methylene blue (MB) on plasmonic aggregates has been ascribed to both direct charge transfer of hot electrons [69] and enhanced plasmon near-fields, which pump energy into molecular adsorbates [57], [91].
1.3.1 Experimental progress exploring the photothermal effect
To address the difficulty of accurately measuring the reaction temperature under illumination, Everitt and Liu developed a methodical experimental approach aimed at achieving identical thermal profiles with and without direct illumination [83]. They studied the plasmon-enhanced CO2 methanation on Rh/TiO2 catalysts in a system comprised of two thermocouples embedded in the catalyst reactor to accurately measure the temperature gradients (Figure 2(c)). They measured the reaction rate under standard direct illumination and under what they called ‘indirect’ illumination. The latter was achieved by using Ti2O3 as a photothermal heater: this material is inert (i.e. is not involved in the CO2 methanation reaction) and converts all the absorbed light into heat, allowing them to realize identical temperature gradients as in the case of the reactor without Ti2O3 (direct illumination), while suppressing the generation of near-fields and hot electrons. The authors found that the combination of thermal and non-thermal effects can synergistically enhance the reaction rates of CO2 methanation on Rh/TiO2 catalysts, and that non-thermal contributions are significant. Interestingly, when adopting a similar approach for the NH3 decomposition on ruthenium catalysis, they found that the reaction was mainly driven by thermal effects [92].
Another rigorous approach to distinguish between these effects was recently proposed by Elias et al. [84]. By developing an annular quartz tube photoreactor with a controllable rate of mass transport and with thermocouples embedded in the catalyst bed, they were able to accurately measure the macroscopic equilibrium temperature and CO oxidation rate on Ag/Al2O3 catalysts in situ. Investigation of the reaction kinetics under external heating and visible light illumination revealed that the rise in equilibrium temperature by itself could not explain the enhanced reaction rate observed during illumination (Figure 2(d)). This remains true even at elevated catalyst loading with larger metal nanocluster sizes (>6 wt%), where photothermal heating is known to be greater and particles are shown to have collective effects [17], [93]. Based on the results, the authors proposed that highly localized effects play a critical role in driving the plasmon-enhanced chemical reactions, including strong near-fields and electronic excitation of reactants from highly localized photothermal heating.
Ultimately, to disentangle these effects, care should be taken while designing the reaction conditions and accurate methodologies should be employed to quantify the co-existence of thermal and non-thermal effects. These can range from simple experimental procedures [53], [94], to more elaborate computations and methodologies [95], [96], [97], [98].
Commonly used techniques to measure the temperature increase in plasmonic systems can be categorized into surface-averaged and single-particle methodologies [99]. The former are beneficial when collective effects are present and the temperature profile is homogenized and include the use of thermocouples, thermoreflectance measurements and thermal cameras. Thermocouples and thermoreflectance measurements translate temperature changes into electric signals or reflectance differences, offering a spatial resolution of approximately 1 µm. In contrast, thermal cameras convert infrared energy into an optical image with a spatial resolution of around 100 µm. Due to their simplicity and minimal impact on the system, these are often relied upon for temperature measurements during catalytic gas-phase reactions [19], [48], [52], [64], [100], [101]. The latter are often used when higher resolution is needed, for instance during the mechanistic investigation of plasmon-mediated chemical reactions via ultra-fast spectroscopy [89] or surface-enhanced Raman scattering [69]. These rely on anti-Stokes thermometry [102], [103], [104] and spectrally measure the anti-Stokes photoluminescence of metallic nanoparticles or probe molecules adsorbed on the surface of plasmonic elements to evaluate the temperature increases of individual nanoparticles in situ. Other techniques with high spatial resolution and single-particle capability are scanning probe microscopies including atomic force and scanning thermal microscopy, where the temperature increase is measured through conductivity changes or using the tip as a thermocouple [105], [106].
Despite the experimental difficulties in distinguishing between non-thermal and thermal effects, careful experimentation has repeatedly shown that the mechanisms responsible for plasmon-assisted catalysis depend on the reaction studied, and most importantly, that the different non-thermal and thermal effects can cooperate synergistically to drive or enhance the catalytic process [9], [51], [107].
1.4 Near-fields or hot electrons?
As previously discussed, non-thermal energy transfer between adsorbed molecules and plasmonic structures can be mediated by optical near-fields and hot carriers and can promote the system to an excited potential energy surface (PES). Both processes benefit from interaction – or hybridization – between metal states and frontier molecular orbitals; however, the strength of hybridization can favour one mechanism or the other. In a recent work, Kazuma et al. employed scanning tunnelling microscopy (STM) to monitor the dissociation reaction of dimethyl disulfide ((CH3S)2) on Ag and Cu surfaces in real-space and real-time [96]. The authors performed measurements under different illumination and bias conditions and found that the reaction yield strongly correlates with the near-field intensity of the optically excited LSPR. They demonstrated that this reaction was driven by direct intramolecular excitation by combining wavelength-dependent reaction yield experiments with finite-difference time-domain simulations to compute the near-field distribution and density functional theory (DFT) to investigate the electronic structure and molecular orbitals of (CH3S)2 adsorbed on metal surfaces. In this case, the strong plasmonic near-fields generated at the nanogaps between the STM tip and the metal surface drive charge excitation from the molecular HOMO–LUMO, as opposed to the transfer of hot electrons from the metal to the molecule. This energy transfer was accessible despite the fact that light with energy below that of the HOMO–LUMO of isolated (CH3S)2 because (CH3S)2 weakly hybridizes with the metal substrates, thus reducing its optical energy gap [108] and promoting the reaction through the pathway in Figure 3(a).
![Figure 3:
Near-fields and hot electrons energy transfer. Subtle difference in reaction pathways driven by direct intramolecular excitation (a) and hot electron transfer to induce a TNI (b). (c) Proposed energy transfer mechanisms involving near-field enhanced excitation of TNIs prepared through charge transfer processes. ψ
i represents the electronic state of the adsorbate. (a–b) Adapted with permission from [96]. Copyright 2018, the American Association for the Advancement of Science (c) Adapted with permission from [110]. Copyright 2020, American Chemical Society.](/document/doi/10.1515/nanoph-2023-0793/asset/graphic/j_nanoph-2023-0793_fig_003.jpg)
Near-fields and hot electrons energy transfer. Subtle difference in reaction pathways driven by direct intramolecular excitation (a) and hot electron transfer to induce a TNI (b). (c) Proposed energy transfer mechanisms involving near-field enhanced excitation of TNIs prepared through charge transfer processes. ψ i represents the electronic state of the adsorbate. (a–b) Adapted with permission from [96]. Copyright 2018, the American Association for the Advancement of Science (c) Adapted with permission from [110]. Copyright 2020, American Chemical Society.
In contrast, the dissociation of strongly hybridized small molecules – such as H2 and O2 [86], [109] – on plasmonic nanoparticles is attributed to hot-electron transfer. Even though the HOMO–LUMO gap of these molecules is accessible only by UV light, these dissociation reactions have been observed to occur under visible-light irradiation on plasmonic catalysts. Density functional theory computations were used to investigate the role and feasibility of hot electron transfer from the metal NPs to the antibonding orbital of these adsorbed molecules and supported the experimental results. These studies demonstrated that high-energy electrons transferred to the antibonding orbitals of H2 and O2, generating TNIs (H2 δ− and O2 δ−), exciting the system to a higher PES and driving the dissociation reaction along the pathway in Figure 3(b).
Recently, an interesting point of view has been proposed in the discussion on plasmon reactions driven by non-thermal effects [110]. Building on evidence from the N-demethylation reaction of methylene blue [57] and knowledge from gas phase dissociation reaction of CO2 [111], Habteyes suggested that near-fields play a more significant role in plasmon-driven photochemical reactions than hot electrons [110]. He takes issue with the current implication in the literature that the TNIs formed by hot carrier transfer undergo an automatic and spontaneous chemical transformation and thus that the plasmon action mechanism is often ascribed to be driven by hot carriers. In contrast, this work proposes that most, if not all, plasmon-mediated reaction are driven by direct excitation of adsorbates and adsorbate–metal complexes by near-fields (Figure 3(c)). He argues that the role of charge transfer is limited to preparing the TNIs: electron transfer results in anionic complexes with the form of S∙[Molecule-L n ] − (where S is the metal surface, L can be a ligand or an environmental molecule and n is an integer), which are stabilized through surface–molecule-environment interactions, after which the reaction proceeds via near-field energy transfer. These newly formed anionic complexes act as intermediate species and could explain the apparent reduction of activation barriers [110]. This mechanism assumes hybridization between the metal NP and reactant species and implies some degree of cooperation between multiple mechanisms.
This was recently confirmed by using Au dimers with varying interparticle distances from 5 nm to 30 nm to monitor the dehalogenation reaction of 4-iodothiophenol (4-ITP) [112]. By using in situ SERS spectroscopy, the authors found that the gap size significantly affects the reaction kinetic. Reducing the separation of the Au dimers from 30 nm to 10 nm resulted in an approximately fourfold increase in the reaction rate. The authors demonstrated that the reaction is predominantly governed by the hot-electrons generated at the gaps, and that stronger near-fields are crucial for efficient hot carrier generation. Similarly, the synergy between strongly confined near-fields and enhanced hot carriers generation has been demonstrated in bimetallic plasmon photocatalysts for the production of hydrogen [113]. This work showed that Au–Pd core-satellite architectures achieve better H2 generation performance compared to core–shell systems, due to the excitation of highly localized and asymmetric near-fields in the gap of core-satellites structures [113].
As the near-fields have a significant impact over the generation of hot carriers through increased photon absorption at the metal surface, balancing the near-fields enhancement and absorption of the plasmonic system is crucial to develop effective photo-activated catalysts [114], [115].
In summary, there is an increasing body of literature that agrees on the synergistic role of non-thermal effects in driving or enhancing chemical reactions. However, their relative weighting is still unclear and likely depends on the system. Detailed theoretical methodologies could help answering this question.
2 Recent progress in atomistic modelling of chemical reactions
2.1 Time-dependent density functional theory
To complement the experimental evidence of plasmon-assisted photocatalysis and to better understand the mechanisms underlying plasmon photocatalysis, considerable effort has been devoted to the study and prediction of excited-state processes. While the optical response of plasmonic structures is accessible with classical electromagnetic simulations [17], [27], [58], investigation of the dynamics of excited plasmon states requires complex quantum approaches. In recent years, significant effort has been devoted in developing theoretical frameworks for the hot carrier generation in plasmonic structures.
Seminal works have framed the current understanding of hot electron generation using different levels of approximation [116], [117], [118], [119]. Govorov et al. developed a simple theoretical framework based on an electron-gas model for various geometries to evaluate the initial energy distribution of hot-carrier generated by plasmon decay [62], [119]. Manjavacas et al. used a jellium approximation to study the plasmon-induced hot carrier process and extended the results to spherical Ag nanoparticles and nanoshells [118]. Atwater’s group have used an ab initio methodology accounting for a detailed electron structure to improve the electronic description of the system [116], [117], while Bernardi et al. [120] combined DFT and electron–phonon calculations to study the energy distribution and scattering processes of the hot carriers generated by SPP in flat surfaces of Au and Ag. All these works contributed to the current knowledge on hot carriers and share similar trends in energy distribution and their dependency on material, band structure and geometry.
These frameworks prompted further advances in atomic-scale modelling of plasmonic hot carrier generation and transfer to molecules or semiconductors, a subject of significant importance in photocatalysis. In series of recent publications, Rossi et al. have developed a fully atomistic methodology to study the plasmon dynamics and electronic excitations of discrete plasmonic systems using a real-time time-dependent DFT (rtTDDFT) approach [41], [121], [122], [123]. The authors have thoroughly analysed the energy and spatial distribution of hot carriers in Ag clusters and showed a pronounced dependence on the size and local structure (Figure 4(a–c)). Upon excitation with a laser pulse at the LSPR frequency, they demonstrated that, as the size of the cluster increases from ∼150 atoms to over 500, the hot carrier distributions are increasingly controlled by interband d-electron transitions and eventually converge to the distributions obtained for flat surfaces [116], [120]. In addition, smaller clusters show a higher population of high-energy carriers, arising from the stronger contributions of sp-states. Interestingly, they found that the spatial distribution of holes is relatively uniform within the cluster with negligible differences between bulk and surface states, while the electrons are more delocalized and reside to a larger extent in the surface region (Figure 4(c)). In particular, lower-coordinated sites (edges and corners) showed larger occupational probability for energetic hot electrons compared to bulk states.
![Figure 4:
Hot carriers distribution and transfer from TDDFT. (a) Hot carrier energy distribution for Ag clusters with different sizes. (b) Probability of hot carrier generation at different atomic sites normalized for the number of atoms for an Ag561 cluster. (c) Spatial distribution of hot holes, hot electrons and hot electrons with energy >1eV for a Ag561 cluster. (d) Molecular PDOS, hot electron distribution and transfer probability as a function of the distance for Au201 + CO on a (111) on-top site (i-iii) and for the same for CO on a corner site (iv-vi). (a–c) Reproduced with permission from ref [41]. Copyright 2020, American Chemical Society. (d) Reproduced with permission from [121]. Copyright 2022, American Chemical Society.](/document/doi/10.1515/nanoph-2023-0793/asset/graphic/j_nanoph-2023-0793_fig_004.jpg)
Hot carriers distribution and transfer from TDDFT. (a) Hot carrier energy distribution for Ag clusters with different sizes. (b) Probability of hot carrier generation at different atomic sites normalized for the number of atoms for an Ag561 cluster. (c) Spatial distribution of hot holes, hot electrons and hot electrons with energy >1eV for a Ag561 cluster. (d) Molecular PDOS, hot electron distribution and transfer probability as a function of the distance for Au201 + CO on a (111) on-top site (i-iii) and for the same for CO on a corner site (iv-vi). (a–c) Reproduced with permission from ref [41]. Copyright 2020, American Chemical Society. (d) Reproduced with permission from [121]. Copyright 2022, American Chemical Society.
These results support the concept of lower-coordinated sites being more catalytically active in heterogeneous catalysis [124], [125] and justify the quest for morphologies with high curvature [58], as these structures support stronger near-field and superior hot electron generation. This rtTDDFT methodology has been used to investigate the direct charge transfer from plasmonic nanoparticles to adsorbed CO molecules and bulk CdSe semiconductors [72], [126], demonstrating plasmon direct hot electron transfer with efficiencies up to 2 % for the CO molecule and 23 % for CdSe. The significant difference in efficiency is due to the higher density of states (DOS) of the CdSe compared to CO.
Interestingly, while it was previously thought that molecules had to be chemisorbed to the metal surface to enable hot carrier transfer [68], a recent rtTDDFT study [121] from the same group has shown that CO molecules can form hybridized energy states with metal nanoclusters at distances up to 6 Å, allowing direct electron transfer (Figure 4(d)). They showed that hot electron transfer depends non-monotonically on the NP–molecule distance and can be effective at long distances, even before a chemical bond can form. Fine tuning the cluster morphology to achieve spectral overlap between optical excitation the energy and the LSPR can lead to direct electron transfer efficiencies up to 8.9 % for a Ag nanocluster and a CO molecule at 2.7 Å from its surface. They further demonstrated that the distribution of hot electrons directly generated on the molecule mirrors the projected density of states (PDOS) for the hybridized metal cluster-molecule system at ground state. The authors point out that the PDOS is a sufficient indicator for a qualitative prediction of the energy distribution of photogenerated hot carriers, suggesting that ground state calculations can be used as a first approximation to quickly screen potential candidates for photocatalysis.
Similar results on the importance of hybridization for electron transfer were found by Le et al. [127] during the investigation of the plasmon-induced CO2 conversion on Al@Cu2O by ground- and excited-state DFT, using the delta self-consistent field (ΔSCF) approximation. This approach allows the calculation of the electronic structure and the energy of excited states by iteratively adjusting the electron density and wavefunction, allowing for predictions of excitation energies and dynamics in molecular systems. They showed that hybridization between CO2 and Cu2O creates new CO2 antibonding states at energies below 2.2 eV, thus accessible by direct plasmonic-excited hot electrons generated by visible light excitation.
The relative contribution of direct and indirect charge transfer process has been recently investigated by Zhang et al. [128] by real-space TDDFT coupled with non-adiabatic Ehrenfest molecular dynamics for CO2 reduction on Ag20 and Ag147 icosahedral clusters. By analysing the time-evolution of Kohn–Sham states, they showed an interplay between direct and indirect electron transfer to unoccupied CO2 levels, which can be tuned by using different laser powers. At lower laser intensity (<250 mJ cm−2), indirect and direct charge transfer mechanisms cooperatively promote CO2 reduction with a relative contribution of about 60 % and 40 %, respectively, while at high laser intensities (>350 mJ cm−2), the direct transfer is highly favoured with a contribution close to 100 %. Furthermore, they showed that larger Ag147 clusters are faster in breaking the C=O bond of adsorbed CO2 compared to Ag20 clusters (40 fs vs 80 fs); this was attributed to an intensified electronic density of states near E F with the increase of the cluster size, which give rise to stronger plasmon field, thereby accelerating the reaction [129].
2.2 Other computational approaches
Although DFT calculations are considered the workhorse of ab initio quantum mechanics methodologies, their results are only as good as the level of approximation used. As such, the results from these studies can often only provide guidelines for understanding the processes investigated [130], [131]. This is because of limitations in properly describing the intricate many-body effects, which are approximated by the electron exchange–correlation (XC) functionals.
In CO2 electroreduction on transition metals, it is widely accepted that the final product distribution depends on the binding energy of CO on the catalyst surface [132], [133]. Formation of CO is considered the rate-limiting step of the CO2 reduction, and the binding energy of this key intermediate is often computed by DFT methods. However, inaccuracies inherent in XC functionals (i.e. the self-interaction error) leads to an erroneous estimation of the CO 2π* orbital energy, which in turn leads to the incorrect predictions of the CO adsorption site and free energy [134]. Instead, more accurate methodologies based on the embedded correlated wavefunction (ECW) theory [135] can be used to account for electron–electron correlation effects and can more accurately describe hybridization and charge transfer processes. While this approach has been used to investigate plasmon-mediated chemical reactions including NH3 decomposition [136], [137] dissociation of small molecules [86], [138] and C–H and C–F bond activation [97], [139], there have been no reports of ECW applied to plasmon photoreduction of CO2.
This theory has been used to revisit the understanding of electrochemical CO2 reduction on copper [134], [140]. By employing the EWC theory, Zhao and colleagues found that the first step in CO reduction on Cu(111) involves *COH instead of the previously thought *CHO, with *COH formation kinetically preferred over *CHO formation by 0.84 eV. This could lead to the reaction happening on a different potential energy surface, possibly leading a different product distribution.
Unfortunately, a major limitation of this methodology is that it is computationally demanding, which does not allow for a facile scale-up. To mitigate this, a recent report by Chen and colleagues [141] demonstrated highly accurate computational CO2 reduction on copper surfaces, benchmarked against several experimental, theoretical and analytical results. This was achieved using a hybrid DFT scheme that combines the doubly hybrid XYG3 XC functional with the periodic generalized gradient approximation and has the advantage of achieving high precision in describing metal–molecules interaction. As this XC builds on top of existing DFT methodologies, it has the potential of being easily implemented and scaled-up for larger systems and could lead to a rapid acceleration of computational chemistry for finite and extended systems in heterogeneous catalysis [142].
Extending these findings to the photoreduction of CO2 and other reactions afforded by plasmonic materials is of pivotal importance in progressing the field of hot carrier science, potentially enabling a predictive comprehension of excited state metal surface reactions.
3 Recent progress in experimental CO2 reduction
Solar-driven conversion of CO2 and H2O into valuable chemical fuels is a very promising approach to address current energy and environmental challenges. Using hydrogen produced by water splitting and CO2 directly captured from the air enables the opportunity to synthesize virtually any valuable hydrocarbon or alcohol.
The challenge in achieving commercially viable production of such valuable fuels is realizing high yields of the desired product in a scalable system and is a critical object of research. In the following sections, we will discuss the most recent advancements on the use of plasmon catalysts for CO2 reduction. We will organize the experimental finding into two categories: in situ measurements, which are primarily used to investigate the reaction mechanisms, and ensemble measurements, which aim to demonstrate optimal performance. In this context, ensemble studies refer to research that adopts macroscopic analytical tools like gas chromatography (GC) and mass spectroscopy (MS) to quantify the product distribution. In the quest to understand the plasmon action mechanisms and engineer efficient plasmonic photocatalysis, many works have used a combination of in situ and ensemble measurements. Consequently, this section will not provide an exhaustive review of all research on plasmon-enhanced photocatalysis, but rather it aims to offer an informative and succinct summary of experimental methodologies for probing reaction pathways and mechanisms, as well as to showcase the significant contributions made by pioneering and cutting-edge research in this field. For a comprehensive introduction to experimental characterization techniques for plasmon-assisted chemistry, we refer the readers to the excellent Reviews [99], [143], [144].
3.1 In situ measurements
In situ techniques enable detailed insights into the structure–activity relationship of catalysts and are commonly used to investigate reaction pathways, intermediates and mechanisms. This is crucial for designing efficient, selective and stable catalytic systems.
3.1.1 Vibrational spectroscopies
Vibrational spectroscopies – including Raman and infrared (IR) spectroscopies – are amongst the most widely used in situ techniques and are used to identify structural information of molecules adsorbed on surfaces. They generate spectra that capture the characteristic vibrations of molecules (fingerprints), enabling non-invasive monitoring of the dynamics of chemical reactions with sub-second resolution. Owing to their high sensitivity and specificity, these techniques are useful to obtain information on possible reaction pathways. Raman spectroscopy is based on inelastic scattering of incident monochromatic light by a molecule and relies on the change of its polarizability. The energy of the scattered light is shifted by an amount corresponding to the vibrational energy of the molecules and this shift provides detailed information about the chemical bonds, molecular structure and composition of the material. The IR and Raman cross sections represent the likelihood of the respective phenomena and are extremely small [145] (10−20 and 10−30 cm−2, respectively) for most molecules; however, they can be greatly enhanced by using resonant metallic structures.
The strong and highly localized near fields generated at the surface of metallic nanoparticles are routinely used to amplify the vibrational Raman signal of molecules, in a technique known as surface-Raman enhanced spectroscopy (SERS) [32], [37]. This technique has been used by Kumari et al. to investigate the dynamics of CO2 reduction and products on Ag NPs under visible light illumination [146]. The authors have used a microfluidic flow cell and performed gas-phase SERS measurements with a 514.5 nm (∼2.4 eV) continuous-wave (CW) laser, matching the LSPR of Ag aggregates, in air and CO2 atmospheres (Figure 5(a, i-ii)). When exposed to air with a 30 % humidity, they observed a broad featureless scattering spectrum that remains steady with the time. Conversely, upon introduction of CO2, a stochastic appearance and disappearance of distinct vibrational fingerprints was observed (Figure 5(a iii–iv)). DFT calculations were used to assign the observed vibrational modes to different chemical species. The product distribution of such species is shown in Figure 5(a-v), indicating the presence of multi-electron reduction species such carbon monoxide (CO) and formic acid (HCOOH), as well as simple alcohols. Importantly, this work identified the very first step of plasmonic CO2 reduction on Ag, by observing the key surface intermediate hydrocarboxyl (HOCO*) originating from the transfer of 1e− from the Ag nanoparticle to CO2. Furthermore, they showed that CO2 is weakly bound to the Ag nanoparticles, at a distance of ∼3.4 Å away from the surface of the Ag nanoparticles. This implies that even over relatively long distances, electron transfer between the metal and the CO2 molecules occurs, in agreement with the theoretical calculations discussed in Section 3.
![Figure 5:
In situ SERS photoreduction of CO2 on Ag clusters. (a-i, ii) Temporal maps showing continuously acquired in situ SERS spectra of air in CO2, respectively. (a-iii, iv) Representative SERS spectra that capture some (CO and HCOO-) of the many products observed during the photocatalysis. (a, v) Amount of observed specie in CO2 photoreduction. (a) Reproduced with permission from [146]. Copyright 2018, American Chemical Society. (b) Reproduced with permission from [147]. Copyright 2021, Springer Nature.](/document/doi/10.1515/nanoph-2023-0793/asset/graphic/j_nanoph-2023-0793_fig_005.jpg)
In situ SERS photoreduction of CO2 on Ag clusters. (a-i, ii) Temporal maps showing continuously acquired in situ SERS spectra of air in CO2, respectively. (a-iii, iv) Representative SERS spectra that capture some (CO and HCOO-) of the many products observed during the photocatalysis. (a, v) Amount of observed specie in CO2 photoreduction. (a) Reproduced with permission from [146]. Copyright 2018, American Chemical Society. (b) Reproduced with permission from [147]. Copyright 2021, Springer Nature.
The same group have used in situ SERS on similar Ag NPs to study the photoreduction of CO2 in the liquid phase, using water as a reaction medium under 514.5 nm CW illumination, in an attempt to extend the product distribution of plasmonic catalysis [147]. They rigorously analysed 42,000 SERS spectra and were able to detect a wide range of products and intermediates, including multi-carbon (C1–C4) species. Out of all the collected spectra, just ∼26 % showed vibrational fingerprints different from the control experiments, and ∼15 % were identified (Figure 5(b, i-ii)). The assignment of the different peaks in the SERS spectra was thoroughly validated through DFT computations and isotope measurements using 13C. The results of these measurements are striking: although Ag NPs yield mostly CO as a product, out of the active spectra, 96.5 % of them show multi-carbon products (Figure 5(b, iii)), leading the authors to infer that multi-electron transfer and C–C coupling is favourable on plasmonic-excited NP surfaces. The relative abundance of the detected species categorized by the number of their constituent carbon atoms, as shown in Figure 5(d, iv), demonstrates the presence of a wide range of highly valuable species, like ethanol, ethylene, acetone, propanol and butanol.
This study marked the first identification of C3 and C4 compounds on Ag-catalysed reactions and holds significant implications for plasmonic CO2 photoreduction; however, it is important to note that the abundance depicted in Figure 5(b, iv) is just an occurrence count and it does not provide a quantitative product distribution. Additionally, the authors used high power density laser (108 Wm−2) to investigate the elementary steps and intermediate species of CO2 reduction in an aqueous environment: this facilitated multi-photon excitation and carrier re-excitation processes, which would result in the generation of highly energetic electron–hole pairs comparable to those generated by UV light. Furthermore, the detected species are only present in trace amounts and do not necessarily survive or contribute to the final product profile. However, showcasing a rich catalogue of C2+ species, this study has given a new perspective of the field of plasmonic photocatalysis with respect to the possibility of converting CO2 into valuable C2+ products.
In situ infrared spectroscopies are complementary techniques to SERS and include Fourier-transform infrared spectroscopy (FTIR), diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) and surface-enhanced infrared absorption spectroscopy (SEIRAS). Contrary to SERS, these rely on absorption of infrared light and depend on the change in the dipole moment of a molecule [148]. Whether or not a vibrational mode is active or inactive for Raman or IR techniques depends on the symmetry of the molecules. As IR absorption is particularly sensitive to heteronuclear molecules and polar bonds this technique is more suitable for analysing liquid-phase catalytic reactions, thanks to its high sensitivity to water and OH groups. These technique allow for in-operando conditions and have been widely adopted to investigate the kinetics of electrochemical reduction of CO2 [149], [150], [151], providing important information on surface adsorbates and intermediates, and are recently being translated to the photocatalysis field [8], [152], [153], [154], [155], [156].
A recent work by Shangguan et al. [154] reported that Au NPs with a plasmon response at ∼540 nm (Figure 6(a)) efficiently reduce CO2 to CO in presence of H2O, showing more than a twofold rate increase in H2O compared to H2. This is a surprising result as generally the photocatalytic conversion of CO2 on metal NP in water requires reducing agents [157], [158], and it is hindered by the competing hydrogen evolution reaction. To gain molecular insights, the authors performed in situ FTIR measurements (Figure 6(b–d)).
![Figure 6:
In-situ monitoring of reaction intermediates with vibrational spectroscopy. (a) Relative contribution of interband and intraband transition in quantum-sized Au NPs. (b) In situ FTIR spectra of the photocatalytic CO2 reduction process occurring with H2O on Au. (c) Time-evolution of the absorbance of key reaction intermediates. (d) FTIR time-steps of the photocatalytic reaction of CO2, showing strongly adsorbed *COOH after 9 h of reaction. Reproduced with permission from [154]. Copyright 2022, Springer Nature.](/document/doi/10.1515/nanoph-2023-0793/asset/graphic/j_nanoph-2023-0793_fig_006.jpg)
In-situ monitoring of reaction intermediates with vibrational spectroscopy. (a) Relative contribution of interband and intraband transition in quantum-sized Au NPs. (b) In situ FTIR spectra of the photocatalytic CO2 reduction process occurring with H2O on Au. (c) Time-evolution of the absorbance of key reaction intermediates. (d) FTIR time-steps of the photocatalytic reaction of CO2, showing strongly adsorbed *COOH after 9 h of reaction. Reproduced with permission from [154]. Copyright 2022, Springer Nature.
A broad peak centred at ∼1650 cm−1 and attributed to H2O appears with the increase of the reaction time, suggesting that H2O was adsorbed on the Au NPs surfaces. Simultaneously, CO2 was activated by the plasmonic NPs, as evidenced by the CO2 − peak at 1698 cm−1. When two or more of these species are in close proximity, they can interact to form monodentate carbonate (m-CO3 2− identifiable from the strong peak at 1380 cm−1 and at 1510 cm−1) and bidentate bicarbonate (peak at 1458 cm−1). As the reaction progresses, further signatures of adsorbed *COOH appear with time (peaks at 1545 and 1180 cm−1), as shown in Figure 6(c). This peak is retained for up to 9 h, indicating that *COOH is strongly adsorbed on the active sites, thereby contributing to the suppression of the HER. The authors supported these results with DFT computations and proved that the activity is induced by surface Au–O species formed from H2O decomposition, which simultaneously optimize the rate-determining steps in the CO2 reduction and H2O oxidation reactions, lowers the energy barriers for the *CO desorption and *OOH formation and facilitates CO and O2 production.
Similar results were found while assessing the photocatalytic performance of plasmonic Au/CdS, Cu/Cu2O octahedrons and Ag, Cu/β-Ga2O3 [159], [160], [161]. In situ DRIFT measurements were employed to assess the intermediate species formed during the CO2 conversion, which led to selective CO production with yields of a few tens of μmol g−1 h−1. These studies identified adsorbed bicarbonate (HCO3 −) and COO− in the wavenumber range of 1350–1520 cm−1 and 1650 cm−1, suggesting that these are important intermediates in the CO2 reduction which can readily accept protons and electrons to form CO.
3.1.2 Electron spectroscopies
As chemical reactions happen at the interfaces, in situ monitoring of the binding energy between reactants and surface atoms offers valuable insights into the chemical state changes occurring during a reaction and driven by electron exchange, helping understanding reaction mechanisms and kinetics. Near-ambient pressure-X-ray photoemission spectroscopy (NAP-XPS) is a powerful technique that can be used to monitor the surface chemistry with unprecedent resolution while the reaction is occurring [162], [163]. A recent example is the assessment of the role of plasmon excitation in Au@CuPd core–shell composites for the reduction of CO2 to CH4 in water environments [164]. Hu and colleagues achieved ∼100 % selectivity toward CH4 under 400 mW cm−2 full-spectrum light illumination using Au nanorods as light-harvesters combined with catalytically active CuPd alloys shells.
Interestingly, upon monochromatic illumination at 800 nm (∼1.55 eV), the authors measured a stable CH4 production rate over 10 cycles with an apparent quantum efficiency of 0.38 %. The results of this research represent the state-of-the-art of CO2 reduction under NIR illumination. To gain insights into this low-energy photon utilization, the authors have used in situ NAP-XPS. The results of these measurements are shown in Figure 7. Upon light irradiation, the binding energy of the Cu 2p3/2 and Pd 3d5/2 peaks decreased (Figure 7(a–d)), indicating the reduction of Cu and Pd species. This was attributed to accumulation of hot electrons generated by Au LSPR relaxation above the Fermi level (E F). Simultaneously, the peaks attributed to adsorbed carbon and gaseous CO2 (Figure 7(e and f)) increase in intensity and present a positive shift, indicating adsorption of CO2 molecules and their conversion process to hydrocarbons. Combined with DFT simulations, the authors concluded that hot electrons accumulated above the E F because of the presence of quasi-isolated trap states enabled by the strong plasmon electric field, effectively extending the lifetime of hot electrons, and increase the probability of electron re-excitation by another low-energy photon.
![Figure 7:
In-situ monitoring of surface chemistry during CO2 methanation. (a, b) In situ NAP-XPS spectra of Cu 2p3/2, (c, d) Pd 3d5/2 and (e, f) C 1s. Panels b, d and f show a top-view contour plot of the highlighted regions in (a, c and e), respectively. Reproduced with permission from [164]. Copyright 2023, Springer Nature.](/document/doi/10.1515/nanoph-2023-0793/asset/graphic/j_nanoph-2023-0793_fig_007.jpg)
In-situ monitoring of surface chemistry during CO2 methanation. (a, b) In situ NAP-XPS spectra of Cu 2p3/2, (c, d) Pd 3d5/2 and (e, f) C 1s. Panels b, d and f show a top-view contour plot of the highlighted regions in (a, c and e), respectively. Reproduced with permission from [164]. Copyright 2023, Springer Nature.
Another in situ characterization methodology is electron spin resonance (ESR), also known as electron paramagnetic resonance. This technique stems from magnetic resonance principles and can detect the transition of unpaired electrons when exposed to a magnetic field [160]. Due to its remarkable ability to detect low concentrations, reaching approximately 10−12 mol, in situ ESR spectroscopy finds its primary application in tracking the formation of radicals on catalyst surfaces during reactions. This, in turn, helps identify potential reaction pathways and mechanisms. ESR has been used to investigate the CO2 reaction mechanisms on plasmonic catalysts [107], [165], [166], demonstrating the involvement of hot electrons in generating TNIs or reactive oxygen vacancies, which promote the activation of CO2.
3.1.3 Imaging technique
Unfortunately, the spectroscopic techniques discussed above do not offer spatial information. In order to visually capture a chemical reaction with nanoscale accuracy and assess the role of near-fields, Wang and colleagues employed in situ environmental scanning transmission electron microscope (ESTEM) coupled with a gas chromatography-mass spectroscopy setup [167]. They exploited the finely focused and highly energetic electron beam to excite LSPRs in Al nanoparticles placed on graphite flakes. This enabled them to observe the endothermic reduction of CO2 to CO by carbon at room temperature, a process commonly known as the reverse Boudouard reaction (CO2(g) + C(s) → 2CO(g)). The process is schematically shown in Figure 8a and b. Under illumination, the plasmon near-fields induced by the Al NPs drive the reaction, leading to the consumption of the graphite flakes underneath, as shown in the TEM images at different time steps in Fig. 8c and d. Through the quantification of etched graphite near the nanoparticles in a CO2 environment, they were able to determine the reaction rate. Simultaneously, they were able to measure both the temperature and spatial pattern of LSPR modes using electron energy loss spectroscopy (EELS). The researchers concluded that a near-field process facilitated by the Al NPs drove the observed reaction of CO2 with carbon. This was further confirmed by measuring the change of the thickness of the graphite flakes before and after (Figure 8(e)) electron beam illumination and comparing it with the simulated field distribution (Figure 8(f)). As expected, they found a correlation between the strength of the coupled near-fields and the amount of etched graphite. These results are in line with previous in situ ESTEM studies on dehydrogenation reaction on plasmonic Au-PdH x , which demonstrated the role of plasmonic near-fields in altering the active site location of hydride nucleation in PdH x [168], [169]. This innovative approach enables precise spatial resolution of LSPR modes distribution on a nanometric scale, as well as the dynamic behaviour of plasmonic nanoparticles with millisecond precision. As a result, its scale-up holds significant promise in the field of photocatalysis.
![Figure 8:
In-situ visualization of CO2 reduction to CO. (a, b) Schematic of Al NPs on a graphite flake, without and with electron beam illumination, respectively. (c, d) Time-resolved TEM images showing etching of graphite in a CO2 partial pressure of ∼50 Pa with a cluster of Al NPs. (e) Carbon depletion map representing the change of carbon thickness after the electron beam illumination. (f) Field distribution of the coupled Al NPs. Reproduced with permission from [167]. Copyright 2021, Springer Nature.](/document/doi/10.1515/nanoph-2023-0793/asset/graphic/j_nanoph-2023-0793_fig_008.jpg)
In-situ visualization of CO2 reduction to CO. (a, b) Schematic of Al NPs on a graphite flake, without and with electron beam illumination, respectively. (c, d) Time-resolved TEM images showing etching of graphite in a CO2 partial pressure of ∼50 Pa with a cluster of Al NPs. (e) Carbon depletion map representing the change of carbon thickness after the electron beam illumination. (f) Field distribution of the coupled Al NPs. Reproduced with permission from [167]. Copyright 2021, Springer Nature.
In summary, in situ measurements have been used in conjunction with theoretical computations to investigate reactions mechanisms, resulting in important insights insights. In doing so, studies have uncovered the role of intermediates, and hot electrons, and provided evidence that plasmonic photocatalysts can help direct the reaction towards certain pathways. However, more work is needed to fully understand how plasmonic catalysts can enable selectivity.
3.2 Ensemble measurements
The overarching goal in this field is to achieve efficient CO2 conversion with high reaction yield to specific products, though a stable and scalable device.
In this section, we will focus on the most recent ensemble experimental results for CO2 reduction and will help the reader to identify trends and limitations of these systems. This section does not aim to be an exhaustive summary, but rather a synthesis of the seminal works and the current state-of-the-art, which have achieved the highest efficiency, selectivity and production of C2+ species. Table 1 provides a summary of various plasmonic catalysts along with their performance and experimental parameters. For detailed reviews on the materials aspect, we refer the reader to the excellent works in ref [15], [170], [171], [172].
Summary of various plasmon photocatalytic systems for CO2 reduction ordered by decreasing product yield.
| Catalyst | Main product | Minor products | Yield main product | Select., % | Illumination condition | Temp., °C | Notes/other | Ref. |
|---|---|---|---|---|---|---|---|---|
| Black Au–Ni | CO | CH4 | 2464 mmol g−1 h−1 | 95 % | Xe lamp (2.77 W cm−2) | 223 | Yield is normalized only to gNi. Gas phase, flow reactor. Gas ratio CO2 : H2 = 10. Isotopic 13C labelling performed | [8] |
| Au–Ag8Cu1 alloy | CO | CH4 | CO 1468.1 mmol g−1 h−1 CH4 398.9 mmol g−1 h−1 | ∼78 % | 300 W Xe lamp (3.7 W cm−2) | 310 | Gas phase + 20 µL of H2O. Batch reactor. 50 % H2/50 % CO2 feed. Selectivity was computed. Isotopic 13C labelling performed | [107] |
| Ni3N | CO | CH4 | 1212 mmol g−1 h−1 | 99 % | Xenon lamp, 400–1100 nm, 3.006 W cm−2 | 199 | Gas phase, flow reactor. Gas ratio CO2 : H2 = 10, 20. Isotopic 13C labelling performed | [173] |
| Au-grafted Ce0.95Ru0.05O2 | CH4 | CO | 473 mmol g−1 h−1 | ∼100 % | Xe lamp (1.6/5.3 W cm−2) | 340 | Gas phase, flow reactor. 4 % H2, 1 % CO2 and 95 % Ar. Isotopic 18O labelling performed | [9] |
| Quantum-sized Au NPs | CO | CH4 | 4730 µmol g−1 h−1 | ∼100 % | LED@420 nm (73 mW cm−2) | 200 | Gas phase + 20 µL of H2O. Batch reactor. Isotopic 13C and H2 18O labelling performed | [154] |
| Au@Pd NPs in UiO-66-NH2 MOF | CO | CH4 | 3737 µmol g−1 h−1 | ∼80 % | 300 W Xe lamp | 150 | Gas phase, batch reactor Gas ratio CO2 : H2 = 3. | [174] |
| AuCu NPs on SrTiO3/TiO2 nanotube | CO | CH2, C2H4, C2H6, C3H6 | 3770 µmol g−1 h−1 | ∼83 % | 300 W Xe lamp | – | Gas phase, batch reactor. Tot hydrocarbons yield 725 μmol g−1 h−1. Isotopic 13C labelling performed | [175] |
| Au@ZIF-67 | CH3OH | C2H5OH | 2500 µmol g−1 h−1 | ∼95 % | Solar simulator (150 mW cm−2) | – | Aqueous solution. Ethanol yield is ∼480 μmol g−1 h−1 | [176] |
| Au–Ag NPs/TIO2 | CO | CH4, C2H4, C2H6, CH3OH, C3H8 | 1813 µmol g−1 h−1 | ∼97 | Vis: Xe lamp (20 mW cm−2) UV: Hg lamp (150 mWcm−2) | – | Gas phase, batch reactor. Gas ratio CO2 : H2 = 1 | [177] |
| Au NPs/TiO2 | CO | CH4, CH3OH, C2H4, C2H6, C3H6, C3H8 | 1237 µmol g−1 h−1 | – | Xe lamp (10 mWcm−2) | 100 | Gas phase, batch reactor. Gas ratio CO2 : H2 = 1 | [178] |
| Au NPs/TiO2 | CO | CH4, C2H4, C2H6, C3H6 | 1223 µmol g−1 h−1 | 98.9 | Vis: solar simulator (100 mWcm−2) UV: Hg lamp (150 mWcm−2) | 100 | Gas phase, flow reactor Gas ratio CO2 : H2 = 1 | [179] |
| Ag NPs/TiO2 NW | CO | CH4, CH3OH | 983 µmol g−1 h−1 | 98 | Hg lamp (20 mW cm−2) | 100 | Gas phase, flow reactor Gas ratio CO2 : H2 = 1 | [180] |
| Au rod@CuPd2 | CH4 | C2H4, C2H6 | 550 µmol g−1 h−1 | ∼100 % | Xe lamp (400 mW cm−2) | 100 | CO2-saturated solution containing Au rod@CuPd2. Isotopic 13C labelling performed | [164] |
| Au/TiO2 | CO | CH4 | 429 µmol g−1 h−1 | 98 % | Solar simulator (1.44 W cm−2) | 150 | Yield is normalized only to gAu. Pressure = 3.6 bar | [181] |
| Al@Cu2O | CO | CH4 | 360 µmol cm−2 h−1 | ∼100 % | Supercontinuum fibre laser (10 W cm−2) | 180 | Gas phase, flow reactor. CO2/H2 feed | [182] |
| Cu–Ru | CH4 | CO | ∼275 µmol g−1 s−1 | >99 % | Supercontinuum fibre laser (19.2 W cm−2) | 750 | Gas phase, flow reactor. Gas ratio CO2 : CH4 = 1. Isotopic 13C labelling performed | [139] |
| Ag/AgClB | CH3CHO | CO, CH4, C2H4 | 209 µmol g−1 h−1 | 96.9 | 500 W Xenon lamp with AM 1.5G filter (100 mW cm−2) | 25 | Liquid phase with NaHCO3 and triethylamine (TEA). Batch reactor. Isotopic 13C labelling performed | [155] |
| Au/m-ZnO | C2H6 | CH4, CO | 27 µmol g−1 h−1 | ∼65 % | 300 W Xe lamp (595 mW) | 25 | Gas phase + 20 µL of H2O. Batch reactor. | [165] |
| Au/TiO2 | CH4 | C2H6, HCHO CH3OH | 15 μmol m−2 h−1a | – | UV @ 254 nm (20 mW cm−2) | 75 | Gas phase + H2O, batch reactor. aYield was computed. | [183] |
| AgCu–TiO2 nanotubes | C2H6 | CH4 | 14.5 µmol g−1 h−1 | 60.7 % | Solar simulator (100 mW cm−2) | 50 | Gas phase + few droplets of H2O. Batch reactor Isotopic 13C labelling performed | [184] |
| Au–Pd on TiO2 | CH4 | C2H4, C2H6, CO | 14.3 µmol g−1 h−1a | 85 %a | 300 W Xe lamp (853 mW cm−2) | 40 | aYield and selectivity are for total hydrocarbon. Gas-phase + 100 µL of H2O. Batch reactor. Isotopic 13C labelling performed | [185] |
| Au NPs + EMIM-BF4 | CH4 | C2H2, C2H4, C3H6, C3H8 | 4.8 NP−1 h−1 | 50 %a | 532 nm laser (1 W cm−2) | 48 | aSelectivity is for C2+ products. Batch reactor. Aqueous solution + EMIM-BF4 Isotopic 13C labelling performed | [4] |
| Au NPs | CH4 | C2H6 | 0.68 NP−1 h−1 | – | Xe lamp (300 mW cm−2) | 47 | Batch reactor. Aqueous solution (water + IPA) | [157] |
| Rh/Al2O3 | CH4 | CO | 0.1–7 µmol cm−2 s−1 | 90–95 % | UV: @365 nm (3 W cm−2) | 300–600 | Gas phase, continuous flow. Gas ratio CO2 : H2 = 1:3.1. Isotopic D2 labelling performed | [10] |
| CH4 | CO | 0.05–0.9 µmol cm−2 s−1 | Blue LED (2.4 W cm−2) | 300–600 | ||||
| Au/TiO2−x | CH4 | C2H6 | 2.5 μmol g−1 h−1 | 60 % | 300W Xe lamp | – | Gas phase + 400 µL of H2O. Batch reactor. | [186] |
| Cu–TiO2 | CH4 | – | 124 ppm cm−2 h−1 | – | AM1.5 illumination (100 mW/cm2) | Gas phase, batch reactor with wet CO2 |
There is a wide variability in terms of reaction yield, product distribution and selectivity, which can be largely attributed to the different experimental conditions (liquid-phase vs gas-phase, broadband vs monochromatic excitation, batch vs flow reaction) resulting from a lack of standard procedures in the field [187]. Variations in the emission spectra of the light source, reaction conditions and reactor configurations have been shown to lead significant changes in the product amounts and distribution for the same reaction and catalyst [187], [188], [189], [190]. There are also discrepancies in the units used to report the yield: while most works report the product amount in moles per gram per hour (µmol g−1 h−1), others normalize to the illuminated area (µmol cm−2 h−1) or use parts-per-million (g−1 h−1), and some use the concept of turnover number. To allow for comparisons between experimental findings, it is essential to report experimental parameters including reactor design, power and emission spectra of light source, gaseous and liquid volume of the reactor, reaction temperature and pressure, amount of catalyst used and its specifications (porosity, active surface area, absorption), sacrificial agents (if present), stability, turnover number and frequency. For these reasons, the field of plasmonic photocatalysis would benefit from the implementation of standard procedures for experimental conditions and data reporting.
Table 1 shows that most reported yields are small, on the order of few tens or hundreds of μmol g−1 h−1, and the studies with the highest yields, in the range of mol g−1 h−1, produce single carbon atom products. Despite this, these works provide fundamental understanding of reaction mechanisms, demonstration of C2+ products and proof of concept of selectivity in CO2 photoreduction.
For instance, the work by Yu et al. demonstrated the kinetically challenging multi-electron, multi-proton, plasmonic-assisted photocatalytic CO2 reduction to methane (CH4) and ethane (C2H6) [157]. Using dispersed Au NPs in water with isopropanol (IPA) as a sacrificial hole scavenger, the authors achieved turnover numbers (TONs) of ∼6.8 and ∼5.6 NP−1 after 10 h of illumination, for CH4 and C2H6, respectively (Figure 9(a, i)). Additionally, they showed that product formation is dependent on the excitation wavelength and intensity (Figure 9(a, ii)): under 532 nm illumination, corresponding to the LSPR of Au NPs, CH4 is the only product and hence the electron harvesting rate is linearly dependent on the laser intensity (Figure 9(a, ii)). However, when illuminating at λ ex = 488 nm, corresponding to the interband transitions of Au, C2H6 was formed when light intensities reached (300 mW cm−2, as shown by the superlinear dependence of the electron harvesting rate on light intensity in Figure 9(a, ii). In line with previous works [191], [192], the authors proposed that hybridization between the metal and molecule states decreases the molecular HOMO–LUMO gap (Figure 9(a, iii–iv)). Simultaneously, CW illumination cathodically polarizes the NP, making it a source of energetic electrons for CO2 activation, thus forming a radical ion intermediate, CO2 •–. This highly active TNI then leads to the formation of CH4. Under high light intensity and interband excitation, more than one electron transfer can take place within the surface residence time of adsorbed CO2, resulting in the simultaneous activation of two CO2 adsorbates, which can couple and promote the formation of C2H6.
![Figure 9:
Plasmon-driven CO2 reduction to C2+ species. (a-i) Turnover number as a function of the reaction time for the CH4 and C2H6 products. (a-ii) Rate of hot electrons harvesting for the two products as a function of the light intensity. (a-iii, iv) Schematic of the plasmon-assisted CO2 reduction mechanism and formation of C1 and C2 products. (b-i, ii) Turnover frequency and selectivity of the produced hydrocarbons as a function of the EMIM-BF4 concentration, respectively. (b-iii, iv) DFT geometries of [EMIM*-CO2] and [H2O–CO2] complexes, respectively. C, H, O and N atoms are grey, white, red and blue, respectively. (b-v, vi, vii) Computed free energy, ∆G, for adding 1e− to CO2, to [EMIM*-CO2] to CO2 in the presence of EMIM+. The free energy of each species is indicated in parentheses. Scale bars are 1 eV in length. (a) Reproduced with permission from ref [157]. Copyright 2018, American Chemical Society. (b) Reproduced with permission from [4]. Copyright 2019, Springer Nature.](/document/doi/10.1515/nanoph-2023-0793/asset/graphic/j_nanoph-2023-0793_fig_009.jpg)
Plasmon-driven CO2 reduction to C2+ species. (a-i) Turnover number as a function of the reaction time for the CH4 and C2H6 products. (a-ii) Rate of hot electrons harvesting for the two products as a function of the light intensity. (a-iii, iv) Schematic of the plasmon-assisted CO2 reduction mechanism and formation of C1 and C2 products. (b-i, ii) Turnover frequency and selectivity of the produced hydrocarbons as a function of the EMIM-BF4 concentration, respectively. (b-iii, iv) DFT geometries of [EMIM*-CO2] and [H2O–CO2] complexes, respectively. C, H, O and N atoms are grey, white, red and blue, respectively. (b-v, vi, vii) Computed free energy, ∆G, for adding 1e− to CO2, to [EMIM*-CO2] to CO2 in the presence of EMIM+. The free energy of each species is indicated in parentheses. Scale bars are 1 eV in length. (a) Reproduced with permission from ref [157]. Copyright 2018, American Chemical Society. (b) Reproduced with permission from [4]. Copyright 2019, Springer Nature.
The same group have also shown the direct synthesis of C1–C3 hydrocarbons using Au NPs illuminated by a green (532 nm) CW laser (Figure 9(b, i-ii)). While the main product is CH4, they also measured ethylene (C2H4), acetylene (C2H2), propane (C3H8) and propene (C3H6), with selectivity for C2+ hydrocarbons of up to 50 %. These results were enabled by using an ionic liquid – 1-ethyl-3-methylimidazolium tetrafluoroborate (EMIM-BF4) – to stabilize charged intermediates and promote electron transfer at the interface between Au NPs and CO2. Results of DFT computations (Figure 9(b, iii–vii)) showed that CO2 strongly interacts with EMIM-BF4, forming a complex [EMIM*-CO2] and binding to the C2 atom of the imidazole ring with an energy of −0.36 eV, much stronger than the interaction of an H2O molecule and CO2 (Figure 9(b, iii–iv)). The ionic liquid also promotes the bending and thus activation of CO2, converging to a OCO angle of 133.7°. Through a comparative analysis (Figure 9(b, v–vii)) of the free energy required to transfer one electron to the different species participating the CO2 reduction (CO2, [EMIM*-CO2] and CO2 in the presence of EMIM+), the authors have proposed that EMIM-BF4 can promote the transfer of photogenerated electrons from the Au NP to adsorbed CO2, which is otherwise a major kinetic bottleneck in the photocatalytic reduction process.
To the best of our knowledge, this is the only work that has obtained C3 products using photocatalysis with pure (only metals) plasmonic materials. Traces of C3 products (<10 μmol g−1 s−1) have been reported when using plasmonic NPs (Au, Ag) as a co-catalysts in hybrid plasmon-semiconductor systems (TiO2, SrTiO3/TiO2) [175], [177], [178]. In these experiments, the proposed mechanism is that hot electrons (holes) generated from the plasmonic NPs are injected in the conduction (valence) band of the semiconductors, improving charge separation, and promoting CO2 reduction and H2 oxidation, respectively. Despite these promising results for the realization of valuable fuels and chemicals from CO2 and (green) H2, the yields and selectivity for C3 compounds are still very low.
Recently, high selectivity towards C2 products and moderate reaction rates were successfully achieved by exploring different combinations of hybrid plasmonic-semiconductors photocatalysts [155], [165], [184], [185]. Zhao et al. [165] used porous ZnO nanosheets decorated with noble metals – Au, Ag and Pd – (Figure 10(a)) and performed a comparative analysis on the CO2 photoreduction performance. The reaction was carried out in the gas-phase and under solar (λ > 320 nm) illumination, using 20 µL of water as a source of protons, without any other sacrificial agent. Interestingly, the choice of plasmon metal drastically influenced the product distribution: Ag achieved 86 % selectivity towards CO with a yield of ∼25 μmol g−1 h−1 (remain product was CH4 at ∼4 μmol g−1 h−1), conversely Pd achieved 85 % selectivity towards CH4 with a yield of ∼18 μmol g−1 h−1 (remain product was CO ∼3.5 μmol g−1 h−1), while Au surprisingly achieved 65 % selectivity towards C2H6 with a production rate of ∼27 μmol g−1 h−1 (remaining products were CH4 and CO, both at ∼20 μmol g−1 h−1). This study demonstrated that CO2 photoreduction with Ag and Pd results only in C1 species, while Au preferentially leads to production of C2H6 which is formed by converting CH4 via a dehydrogenative coupling mechanism (2CH4 → C2H6 + H2) [193]. The reaction was further investigated by DFT calculations, spin trapping electron paramagnetic spectroscopy and photo-electrochemical measurements. The authors suggested that the introduction of plasmon materials have a twofold purpose: (i) the plasmon near-fields couple and enhance the spontaneously generated inner electric field formed between stacked ZnO nanosheets, resulting in a more efficient charge separation and increased carrier lifetime, and (ii) enable plasmon resonant energy transfer across the ZnO bandgap via interband transition.
![Figure 10:
CO2 photoreduction to C2+ species with hybrid plasmon-semiconductor materials. (a) Schematic representation of Au/ZnO nanosheets for CO2 photoreduction and (b) product evolution under light illumination. (c) Absorptance of TiO2 nanotubes arrays decorated with AgCu NPs along with its schematic representation and field enhancement at the LSPR. (d) Product distribution of TiO2 nanotube arrays decorated with different plasmonic NPs. (e) Photocatalytic CH3CHO production activity and selectivity of the Janus Ag/AgCl0.79Br0.21 nanostructures as a function of the reaction time. (f) Effect of light irradiation on the product yield and selectivity. (a–b) Reproduced with permission from [165]. Copyrights 2019, Elsevier. (c–d) Reproduced with permission from [184]. Copyright 2021, American Chemical Society (e–f) Reproduced with permission from [155]. Copyrights 2021, Royal Society of Chemistry.](/document/doi/10.1515/nanoph-2023-0793/asset/graphic/j_nanoph-2023-0793_fig_010.jpg)
CO2 photoreduction to C2+ species with hybrid plasmon-semiconductor materials. (a) Schematic representation of Au/ZnO nanosheets for CO2 photoreduction and (b) product evolution under light illumination. (c) Absorptance of TiO2 nanotubes arrays decorated with AgCu NPs along with its schematic representation and field enhancement at the LSPR. (d) Product distribution of TiO2 nanotube arrays decorated with different plasmonic NPs. (e) Photocatalytic CH3CHO production activity and selectivity of the Janus Ag/AgCl0.79Br0.21 nanostructures as a function of the reaction time. (f) Effect of light irradiation on the product yield and selectivity. (a–b) Reproduced with permission from [165]. Copyrights 2019, Elsevier. (c–d) Reproduced with permission from [184]. Copyright 2021, American Chemical Society (e–f) Reproduced with permission from [155]. Copyrights 2021, Royal Society of Chemistry.
The effect of different metals on the production of C2 species was also investigated by Vahidzadeh et al. [184]. Here, the authors tested densely packed TiO2 nanotubes arrays (TNTA) decorated with Ag, Cu and AgCu alloys NPs towards the visible (AM 1.5G 1-sun illumination) gas-phase photoreduction of CO2 without using any sacrificial agent or hole scavengers (Figure 10(c)). They found that the addition of large sized (80–200 nm) NPs increases both the efficiency and selectivity of CO2 photoreduction. As shown in Figure 10(d), the AgCu-TNTA exhibited higher product yields (14.5 μmol g−1 h−1 of C2H6 and 9.38 μmol g−1 h−1 of CH4) and higher selectivity (60.7 % towards ethane) for C2 products compared to TNTAs decorated with monometallic Ag or Cu NPs. The authors attributed the superior performance towards C2H6 to cooperative effects of (i) the high plasmonic hot-spots density, (ii) the asymmetric charge distribution generated from closely packed NPs which decreases adsorbate–adsorbate repulsion and improves C–C coupling and (iii) increased lifetime of hot electrons injected over the Schottky barrier formed at the metal–semiconductor interface, although a rigorous mechanistic study was missing. As hot electrons transfer is widely accepted to be more efficient in small nanoparticles [17], [62], [194], the high performance achieved by using large plasmonic NPs (>80 nm) suggests that in these systems near-field enhanced mechanisms are more significant than hot electrons transfer.
While most of the studies have detected gas-phase products, converting CO2 into liquid (or solid) compounds holds significant practical importance because of their higher energy density and ease of storage and transportation [195], [196]. In a recent study [155], broadband Janus silver/ternary silver halide (Ag/AgClBr) nanostructures were shown to achieve a record-high 96.9 % selectivity towards acetaldehyde (CH3CHO) with a generation rate of 209 μmol g−1 h−1 under solar illumination, while just traces of other species (CO, CH4 and C2H4) were observed (Figure 10(e)). The photocatalysts were dispersed into an aqueous solution containing 0.1 M NaHCO3 and triethylamine as a hole scavenger. This is a surprising result as the formation of CH3CHO is an energy intensive reaction, which requires transfer of 10 electrons. The authors also performed several control experiments, including the 13CO2 isotope labelling, to confirm that acetaldehyde was effectively being produced by the CO2 reduction on the catalyst. To investigate the mechanism, the authors performed the reaction under full-solar illumination (plasmon-on state in Figure 10(f)) and λ ex > 420 nm, to prevent the excitation of the Ag LSPR at 350 nm (plasmon-off state in Figure 10(f)). Upon solar illumination (plasmon-on), the authors detected more than a threefold increase in the reaction yield along with an improved selectivity compared to the plasmon-off state. They attributed the increase in performance to a synergistic effect of enhanced near-fields and hot electron transfer, which promotes the system to an excited potential energy surface, improving selectivity and CH3CHO production rate.
In order to improve the productivity of plasmon-enhanced CO2 conversion, an inspiring strategy is to use broadband absorbers. Recently, dendritic plasmonic colloidosomes of Au loaded with nickel, DPC-C4-Ni, (Figure 11(a, i)) achieved extremely high performances for the gas-phase CO2 hydrogenation reaction [8]. The authors have taken advantage of the broadband absorption and high surface area of the composite structure, along with the enhanced catalytic activity of Ni, to achieve a record-high CO production rate of 2464 ± 40 mmol gNi −1 h−1 with 95 % selectivity (Figure 11, (ii–iii)) and stability up to 100 h. The reaction was carried out in a flow reactor at atmospheric pressure and without external heating. It is worth noting that such high rate is atypical for photocatalysis – the amount of products is generally in the tens or hundreds of μmol g−1 h−1 (see Table 1) – and it is comparable with reaction rates obtained from electrocatalysis. Upon illumination, the plasmon relaxation process described in Section 2 leads to a catalyst surface temperature of 223 °C. Interestingly, a comparison of the reaction yield under illumination and in the dark (with external heat provided to reach the same temperature as under illumination) revealed a 9-fold increase in activity in light as compared to the dark and demonstrating the importance of non-thermal effects. In situ DRIFT spectroscopy showed that CO is weakly bonded to the active Ni sites, inferring that its desorption is efficient, thus restricting further hydrogenation to CH4 and leading to ∼95 % CO selectivity. Meanwhile, findings from kinetic isotope effect and ultrafast transient absorption spectroscopy demonstrated that hot electrons generated from the gold NPs are indirectly transferred to the catalytically active Ni sites and then participate in the reaction, similarly to antenna-reactor systems [101], [139], [164].
![Figure 11:
High performing CO2 to C1 products photocatalysis. (a-i) schematic of the synthesis of Ni-laden black-gold dendritic colloidosomes. (a-ii) Plasmonic CO2 hydrogenation reaction yield upon multiple light-on light-off cycles, without any external heating. (a-iii) Products selectivity measured at different light intensities. (b-i) Absorption profiles of different stacked plasmonic metamaterials. (b-ii) Production rate of CO and CH4 for plasmonic CO2 hydrogenation and (b-iii) comparison between reaction rates obtained in dark and light conditions for the different metamaterials. (c-i) Schematic of the cooperative photochemical and photothermal effects in the plasmon-enhanced CO2 methanation process (c-ii). CO2 conversion, CH4 selectivity and production rate of the CO2 methanation on Au0.1/Ce0.95Ru0.05O2 as a function of the gas hourly space velocity (GHSV) using a concentrated reactant gas. (a) Reproduced with permission from [8]. Copyright 2023, American Chemical Society. (b) Reproduced with permission from [107]. Copyright 2022, Wiley-VCH GmbH. (c) Reproduced with permission from [9]. Copyright 2023, Springer Nature.](/document/doi/10.1515/nanoph-2023-0793/asset/graphic/j_nanoph-2023-0793_fig_011.jpg)
High performing CO2 to C1 products photocatalysis. (a-i) schematic of the synthesis of Ni-laden black-gold dendritic colloidosomes. (a-ii) Plasmonic CO2 hydrogenation reaction yield upon multiple light-on light-off cycles, without any external heating. (a-iii) Products selectivity measured at different light intensities. (b-i) Absorption profiles of different stacked plasmonic metamaterials. (b-ii) Production rate of CO and CH4 for plasmonic CO2 hydrogenation and (b-iii) comparison between reaction rates obtained in dark and light conditions for the different metamaterials. (c-i) Schematic of the cooperative photochemical and photothermal effects in the plasmon-enhanced CO2 methanation process (c-ii). CO2 conversion, CH4 selectivity and production rate of the CO2 methanation on Au0.1/Ce0.95Ru0.05O2 as a function of the gas hourly space velocity (GHSV) using a concentrated reactant gas. (a) Reproduced with permission from [8]. Copyright 2023, American Chemical Society. (b) Reproduced with permission from [107]. Copyright 2022, Wiley-VCH GmbH. (c) Reproduced with permission from [9]. Copyright 2023, Springer Nature.
Reaction rates of the same order of magnitude were also achieved by using a novel non-noble metal plasmonic material for CO formation. By exploiting the plasmonic properties of hydroxy-terminated Ni3N nanosheets, Singh et al. [173] demonstrated efficient CO2 hydrogenation reaction, obtaining high CO production rate of 1212 mmol g−1 h−1 with ∼99 % CO selectivity, under visible illumination, in a flow reactor and without any sacrificial agent. Nickel is widely used in electrocatalysis to achieve efficient water splitting [197], [198], and these results shown that it also has high catalytic activity and affinity towards CO production, suggesting that Ni and metal nitrides could be used to develop efficient plasmonic photocatalysts.
The CO2 hydrogenation reaction was also recently studied by Shao et al. [107] with broadband periodic metamaterials. Contrary to the previous works, this research intentionally exploited the strongly localized near-fields and hot electrons generated from the plasmonic elements to achieve intense localized temperature increase on the device to enable the photo-thermal CO2 conversion. The authors used a multistep template-assisted colloidal lithography technique to fabricate a plasmonic metamaterial consisting of stacked Aufilm–SiO2–Aupatterned-SiO2-Ag8Cu (ASA-c-Ag8Cu) layered structure. Here, the Aupatterned consists of tapered triangular arrays and the Au film acts as a rear mirror and is necessary to achieve >80 % absorption across the vis-NIR (Figure 11(b, i)) by increasing internal light scattering, SiO2 improved thermal stability by preventing heat deformation of the Aupatterned and finally Ag8Cu provides the catalytically active site. Thanks to the enhanced light absorption, localized near-fields and accumulation of hot electrons, this complex structure achieves surface temperatures as high as 310 °C under full-spectrum light illumination at 3.7 W cm−2. This led to reaction yields of 398.9 and 1468.1 mmol g−1 h−1 for CH4 and CO, respectively. Figure 11(b, ii–iii) reports these production rates in units of mmol m−2 h−2. As shown in Figure 11(b-iii), the reaction rates in the dark at the same temperature are substantially lower than those under light irradiation, indicating the importance of plasmonic excitation, in line with previous works [8], [173]. Further results obtained from electrodynamic simulations and in situ XPS and in situ electron paramagnetic resonance (EPR), suggested that all the plasmon relaxation processes synergistically contribute to promoting CO2 activation and reducing the energy barrier for CO2 hydrogenation. Strong localized near-fields promote hot electron generation and transfer from Au to the catalytically active Ag8Cu and synergistically enhance the photothermal effect, leading to higher catalytic activities.
A similar strategy was recently used to promote efficient photothermal CO2 methanation on Au/Ce0.95Ru0.05O2 solid-solution catalyst [9]. By utilizing visible-near-infrared light irradiation (350–2500 nm, 5.3 W cm−2) without any supplementary heating, this catalyst demonstrated a high CH4 formation rate of 473 mmol g−1 h−1 with ∼100 % selectivity and single-pass CO2 conversion rate of ∼75 % (Figure 10), surpassing the activity of conventional bandgap-excitation photocatalysts and approaching the thermodynamic limit. This was accomplished in a continuous flow reactor aimed at simulating the industrial CO2 methanation process, by using the same gas composition (72 % H2, 18 % CO2, balance Ar) and similar gas hourly space velocity (GHSV). The latter is an important parameter in CO2 methanation studies as it affects the reaction rate and selectivity of the process and represent the volumetric flow rate of the gas feed per unit volume of the catalyst bed per hour. Upon careful investigation, the authors found that the reaction was initiated by photothermal effects and enhanced by the plasmon excitation. As shown in Figure 11(c-ii, iii), upon absorption of vis-NIR light Ce0.95Ru0.05O2 generates heat, which increases the catalysts surface temperature up to ∼340 °C, while visible excitation of the LSPR generates hot electrons, which are then injected into Ce0.95Ru0.05O2 creating surface oxygen vacancies (VO) near the ruthenium sites. These Ru-VO are the catalytically active centres and accelerate the CO2 methanation process, by facilitating CO2 adsorption and H2 dissociation. The VO exhibit similarities to transient negative ions (TNIs), as both involve charged sites affecting chemical reactions; however, TNIs typically refer to short-lived species that play a role in reaction intermediates, while VO are structural defects influencing catalytic behaviour [199]. Despite the fact that the surface temperature reaches up to ∼340 °C, the bulk temperature of the system is just ∼50 °C, which is notably lower than the conditions required for industrial reactors (200–550 °C with pressures of 1–100 bar). Along with the high reaction rate and excellent selectivity, this study indicates the potential advantages of the LSPR-enhanced photothermal CO2 methanation reaction over conventional energy-intensive thermal systems which, if applied at industrial scale, could offer a clean and sustainable solution for storing intermittent renewable energy [200].
4 Outlook and perspective
To conclude, this review provides a comprehensive overview of the recent advancements in plasmonic photocatalysis for CO2 reduction, focusing on the key mechanisms, challenges and trends. We delved into the fundamental principles underlying the interactions between plasmonic nanoparticles, light and molecular species, elucidating how these interactions can be harnessed to drive efficient energy transfer processes during heterogeneous photocatalysis. We then explored the various experimental strategies that have been used to differentiate between the different competing and/or cooperating effects. Lastly, we summarized the latest cutting-edge research efforts focused on the mechanistic understanding of reaction pathways as well as at achieving high performance in CO2 reduction. We hope to have provided factual demonstration that more often than not, synergistic plasmonic energy transfer mechanisms promote the reaction and increase the efficiency. Near-fields can significantly increase molecular absorption; hot carrier transfer can initiate reactions and plasmonic photothermal heat generation can increase catalytic activity, all of which can result in enhanced photochemical reactions. Throughout our examination, it has become evident that plasmonic photocatalysis holds great promise in addressing the pressing global challenges of converting CO2 into valuable sources of clean energy. However, as we navigate within this field, several challenges and questions persist, including the optimization of plasmonic catalysts, the scalability of these processes and the long-term stability of plasmonic nanomaterials. The path forward will require interdisciplinary collaboration, innovation in materials design and the integration of cutting-edge techniques, such as machine learning, to unlock the full potential of plasmonic photocatalysis.
Recently, the emergence of the new sub-field of vibropolaronic chemistry [201] has shown that strong coupling of molecules with optical modes – where photons are exchanged faster than competing dissipative processes – can lead to new hybrid light–matter polaritonic states, which can modify the potential energy surface. Coupling of vibrational modes with the vacuum field of Fabry–Perot cavity or optical cavity modes [44], [202] can thus provide a new degree of freedom in modulating chemical reactions, thus opening a new avenue for the field of heterogeneous catalyst.
Additionally, understanding the role of hot holes is essential for optimizing solar-driven CO2 reduction strategies and to improve the activity for C2+ species. Better utilization of hot holes could help provide additional protons necessary to produce value-added hydrocarbons and alcohols and, therefore, improve yields. In particular, exploring efficient metal–semiconductor interfaces could improve charge separation and extend the lifetime of photogenerated hot electrons and holes, ultimately increasing reaction rates and selectivity.
Simultaneously, the integration of machine learning algorithms in the field of heterogeneous photocatalysis [203], [204] could help boost the field. The drastic and exponential growth of these approaches, along with their ability in dealing with big data, holds great promise for propelling the field of plasmonic and heterogeneous photocatalysis into new frontiers of discovery and application. By exploiting the power of data-driven algorithms, it is expected that there will be an important shift from DFT methodologies to machine learning approaches, which could enable to gain deeper insights into complex reaction mechanisms, accelerate materials discovery and optimize reaction conditions.
Leveraging the remarkable reaction yields and selectivity towards C1 species achieved by broadband plasmonic materials coupled with semiconductors or other catalytically active species, we anticipate that broadband and highly absorptive plasmon catalysts will play a pivotal role in harnessing and converting electromagnetic energy from sunlight towards the production of valuable C2+ species with high yield and selectivity. As we stand on the threshold of a new era in materials science and catalysis, the marriage of plasmonic photocatalysis with surface science, advances in characterization techniques and computational methodologies offers an exciting and dynamic path forward, promising breakthroughs that may revolutionize energy conversion, environmental remediation and other critical areas of science and technology.
Funding source: Meta Active – International Research Training Group
Award Identifier / Grant number: IRTG 2675
-
Research funding: F.J.B. gratefully acknowledges the support of the Meta Active – International Research Training Group (IRTG 2675).
-
Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.
-
Conflict of interest: Authors state no conflicts of interest.
-
Informed consent: Informed consent was obtained from all individuals included in this study.
-
Ethical approval: The conducted research is not related to either human or animals use.
-
Data availability: Data sharing is not applicable to this article as no datasets were generated or analysed during the current study.
References
[1] H. Ritchie, P. Rosado, and M. Roser, OurWorldinData.org, CO2 and Greenhouse Gas Emissions, OurWorldInData.org, 2023. Available at: https://ourworldindata.org/co2-and-greenhouse-gas-emissions.Search in Google Scholar
[2] S. N. Habisreutinger, L. Schmidt-Mende, and J. K. Stolarczyk, “Photocatalytic reduction of CO2 on TiO2 and other semiconductors,” Angew Chem. Int. Ed. Engl., vol. 52, no. 29, pp. 7372–7408, 2013. https://doi.org/10.1002/anie.201207199.Search in Google Scholar PubMed
[3] J. Wu, Y. Huang, W. Ye, and Y. Li, “CO2 reduction: from the electrochemical to photochemical approach,” Adv. Sci., vol. 4, no. 11, p. 1700194, 2017. https://doi.org/10.1002/advs.201700194.Search in Google Scholar PubMed PubMed Central
[4] S. Yu and P. K. Jain, “Plasmonic photosynthesis of C(1)-C(3) hydrocarbons from carbon dioxide assisted by an ionic liquid,” Nat. Commun., vol. 10, no. 1, p. 2022, 2019. https://doi.org/10.1038/s41467-019-10084-5.Search in Google Scholar PubMed PubMed Central
[5] Y. Kim, J. G. Smith, and P. K. Jain, “Harvesting multiple electron–hole pairs generated through plasmonic excitation of Au nanoparticles,” Nat. Chem., vol. 10, no. 7, pp. 763–769, 2018. https://doi.org/10.1038/s41557-018-0054-3.Search in Google Scholar PubMed
[6] G. Kim and W. Choi, “Charge-transfer surface complex of EDTA-TiO2 and its effect on photocatalysis under visible light,” Appl. Catal., B, vol. 100, no. 1, pp. 77–83, 2010. https://doi.org/10.1016/j.apcatb.2010.07.014.Search in Google Scholar
[7] Y. Wang, E. Chen, and J. Tang, “Insight on reaction pathways of photocatalytic CO2 conversion,” ACS Catal., vol. 12, no. 12, pp. 7300–7316, 2022. https://doi.org/10.1021/acscatal.2c01012.Search in Google Scholar PubMed PubMed Central
[8] R. Verma, R. Belgamwar, P. Chatterjee, R. Bericat-Vadell, J. Sa, and V. Polshettiwar, “Nickel-laden dendritic plasmonic colloidosomes of black gold: forced plasmon mediated photocatalytic CO2 hydrogenation,” ACS Nano, vol. 17, no. 5, pp. 4526–4538, 2023. https://doi.org/10.1021/acsnano.2c10470.Search in Google Scholar PubMed
[9] H. Jiang, et al.., “Light-driven CO2 methanation over Au-grafted Ce0.95Ru0.05O2 solid-solution catalysts with activities approaching the thermodynamic limit,” Nat. Catal., vol. 6, no. 6, pp. 519–530, 2023. https://doi.org/10.1038/s41929-023-00970-z.Search in Google Scholar
[10] X. 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
[11] 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
[12] Y. Yuan, 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
[13] E. Cortes, “Light-activated catalysts point the way to sustainable chemistry,” Nature, vol. 614, pp. 230–232, 2023. https://doi.org/10.1038/d41586-023-00239-2.Search in Google Scholar PubMed
[14] C. Zhan, M. Moskovits, and Z.-Q. Tian, “Recent progress and prospects in plasmon-mediated chemical reaction,” Matter, vol. 3, no. 1, pp. 42–56, 2020. https://doi.org/10.1016/j.matt.2020.03.019.Search in Google Scholar
[15] S. Ezendam, et al.., “Hybrid plasmonic nanomaterials for hydrogen generation and carbon dioxide reduction,” ACS Energy Lett., vol. 7, no. 2, pp. 778–815, 2022. https://doi.org/10.1021/acsenergylett.1c02241.Search in Google Scholar PubMed PubMed Central
[16] E. Cortés, 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
[17] Z. Fusco, K. Catchpole, and F. J. Beck, “Investigation of the mechanisms of plasmon-mediated photocatalysis: synergistic contribution of near-field and charge transfer effects,” J. Mater. Chem. C, vol. 10, no. 19, pp. 7511–7524, 2022. https://doi.org/10.1039/d2tc00491g.Search in Google Scholar
[18] J. Gargiulo, R. Berté, Y. Li, S. A. Maier, and E. Cortés, “From optical to chemical hot spots in plasmonics,” Acc. Chem. Res., vol. 52, no. 9, pp. 2525–2535, 2019. https://doi.org/10.1021/acs.accounts.9b00234.Search in Google Scholar PubMed
[19] L. Mascaretti and A. Naldoni, “Hot electron and thermal effects in plasmonic photocatalysis,” J. Appl. Phys., vol. 128, no. 4, pp. 041101, 2020. https://doi.org/10.1063/5.0013945.Search in Google Scholar
[20] C. Zhang, F. Jia, Z. Li, X. Huang, and G. Lu, “Plasmon-generated hot holes for chemical reactions,” Nano Res., vol. 13, no. 12, pp. 3183–3197, 2020. https://doi.org/10.1007/s12274-020-3031-2.Search in Google Scholar
[21] J. Bian, Z. Zhang, Y. Liu, E. Chen, J. Tang, and L. Jing, “Strategies and reaction systems for solar-driven CO2 reduction by water,” Carb. Neutr., vol. 1, no. 1, p. 5, 2022. https://doi.org/10.1007/s43979-022-00006-8.Search in Google Scholar
[22] J. Li and M. M. Waegele, “Advances in understanding the role of surface hole formation in heterogeneous water oxidation,” Curr. Opin. Electrochem., vol. 33, p. 100932, 2022. https://doi.org/10.1016/j.coelec.2021.100932.Search in Google Scholar
[23] A. S. Urban, et al.., “Three-dimensional plasmonic nanoclusters,” Nano Lett., vol. 13, no. 9, pp. 4399–4403, 2013. https://doi.org/10.1021/nl402231z.Search in Google Scholar PubMed
[24] G. V. Naik, V. M. Shalaev, and A. Boltasseva, “Alternative plasmonic materials: beyond gold and silver,” Adv. Mater., vol. 25, no. 24, pp. 3264–3294, 2013. https://doi.org/10.1002/adma.201205076.Search in Google Scholar PubMed
[25] N. I. Zheludev and Y. S. Kivshar, “From metamaterials to metadevices,” Nat. Mater., vol. 11, no. 11, pp. 917–924, 2012. https://doi.org/10.1038/nmat3431.Search in Google Scholar PubMed
[26] T. Tanaka and A. Ishikawa, “Towards three-dimensional optical metamaterials,” Nano Convergence, vol. 4, no. 1, p. 34, 2017. https://doi.org/10.1186/s40580-017-0129-7.Search in Google Scholar PubMed PubMed Central
[27] Z. Fusco, et al.., “Photonic fractal metamaterials: a metal–semiconductor platform with enhanced volatile-compound sensing performance,” Adv. Mater., vol. 32, no. 50, p. 2002471, 2020. https://doi.org/10.1002/adma.202002471.Search in Google Scholar PubMed
[28] R. G. Hobbs, W. P. Putnam, A. Fallahi, Y. Yang, F. X. Kärtner, and K. K. Berggren, “Mapping photoemission and hot-electron emission from plasmonic nanoantennas,” Nano Lett., vol. 17, no. 10, pp. 6069–6076, 2017. https://doi.org/10.1021/acs.nanolett.7b02495.Search in Google Scholar PubMed
[29] M. B. Ross, J. C. Ku, B. Lee, C. A. Mirkin, and G. C. Schatz, “Plasmonic metallurgy enabled by DNA,” Adv. Mater., vol. 28, no. 14, pp. 2790–2794, 2016. https://doi.org/10.1002/adma.201505806.Search in Google Scholar PubMed
[30] S. Unser, I. Bruzas, J. He, and L. Sagle, “Localized surface plasmon resonance biosensing: current challenges and approaches,” Sensors, vol. 15, no. 7, pp. 15684–15716, 2015. https://doi.org/10.3390/s150715684.Search in Google Scholar PubMed PubMed Central
[31] S. A. Maier, Plasmonics: Fundamentals And Applications, vol. 1, New York, NY, Springer, 2007, p. 224, XXVI.10.1007/0-387-37825-1Search in Google Scholar
[32] P. L. Stiles, J. A. Dieringer, N. C. Shah, and R. P. Van Duyne, “Surface-enhanced Raman spectroscopy,” Annu. Rev. Anal. Chem., vol. 1, no. 1, pp. 601–626, 2008. https://doi.org/10.1146/annurev.anchem.1.031207.112814.Search in Google Scholar PubMed
[33] K. Chen and H. Wang, “Plasmon-driven photocatalytic molecular transformations on metallic nanostructure surfaces: mechanistic insights gained from plasmon-enhanced Raman spectroscopy,” Mol. Syst. Des. Eng., vol. 6, no. 4, pp. 250–280, 2021. https://doi.org/10.1039/d1me00016k.Search in Google Scholar
[34] 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
[35] V. Jain, R. K. Kashyap, and P. P. Pillai, “Plasmonic photocatalysis: activating chemical bonds through light and plasmon,” Adv. Opt. Mater., vol. 10, no. 15, p. 2200463, 2022. https://doi.org/10.1002/adom.202200463.Search in Google Scholar
[36] A. J. Haes, S. Zou, G. C. Schatz, and R. P. Van Duyne, “Nanoscale optical biosensor: short range distance dependence of the localized surface plasmon resonance of noble metal nanoparticles,” J. Phys. Chem. B, vol. 108, no. 22, pp. 6961–6968, 2004. https://doi.org/10.1021/jp036261n.Search in Google Scholar
[37] B. Sharma, R. R. Frontiera, A. I. Henry, E. Ringe, and R. P. Van Duyne, “SERS: materials, applications, and the future,” Mater. Today, vol. 15, nos. 1–2, pp. 16–25, 2012. https://doi.org/10.1016/s1369-7021(12)70017-2.Search in Google Scholar
[38] Z. Fusco, R. Bo, Y. Wang, N. Motta, H. Chen, and A. Tricoli, “Self-assembly of Au nano-islands with tuneable organized disorder for highly sensitive SERS,” J. Mater. Chem. C, vol. 7, no. 21, pp. 6308–6316, 2019. https://doi.org/10.1039/c9tc01231a.Search in Google Scholar
[39] J. B. Khurgin, A. Petrov, M. Eich, and A. V. Uskov, “Direct plasmonic excitation of the hybridized surface states in metal nanoparticles,” ACS Photonics, vol. 8, no. 7, pp. 2041–2049, 2021. https://doi.org/10.1021/acsphotonics.1c00167.Search in Google Scholar
[40] G. Tagliabue, et al.., “Quantifying the role of surface plasmon excitation and hot carrier transport in plasmonic devices,” Nat. Commun., vol. 9, no. 1, p. 3394, 2018. https://doi.org/10.1038/s41467-018-05968-x.Search in Google Scholar PubMed PubMed Central
[41] 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
[42] J. S. DuChene, G. Tagliabue, A. J. Welch, W. H. Cheng, and H. A. Atwater, “Hot hole collection and photoelectrochemical CO(2) 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
[43] B. Jeon, C. Lee, and J. Y. Park, “Electronic control of hot electron transport using modified Schottky barriers in metal-semiconductor nanodiodes,” ACS Appl. Mater. Interfaces, vol. 13, no. 7, pp. 9252–9259, 2021. https://doi.org/10.1021/acsami.0c22108.Search in Google Scholar PubMed
[44] S. Zhao, Z. Fusco, and F. J. Beck, “Strong and tunable absorption in coupled nanoparticle–cavity systems for plasmonically enhanced hot electron devices,” Optica, vol. 9, no. 9, pp. 1084–1093, 2022. https://doi.org/10.1364/optica.465740.Search in Google Scholar
[45] S. Zhao, Y. Yin, J. Peng, Y. Wu, G. G. Andersson, and F. J. Beck, “The importance of Schottky barrier height in plasmonically enhanced hot-electron devices,” Adv. Opt. Mater., vol. 9, no. 3, p. 2001121, 2021. https://doi.org/10.1002/adom.202001121.Search in Google Scholar
[46] J. Khurgin, A. Y. Bykov, and A. V. Zayats, “Hot-electron dynamics in plasmonic nanostructures,” 2023, arXiv:2302.10247.Search in Google Scholar
[47] P. B. Allen, “Theory of thermal relaxation of electrons in metals,” Phys. Rev. Lett., vol. 59, no. 13, pp. 1460–1463, 1987. https://doi.org/10.1103/physrevlett.59.1460.Search in Google Scholar PubMed
[48] L. Mascaretti, et al.., “Challenges and prospects of plasmonic metasurfaces for photothermal catalysis,” Nanophotonics, vol. 11, no. 13, pp. 3035–3056, 2022. https://doi.org/10.1515/nanoph-2022-0073.Search in Google Scholar
[49] X. Wan, et al.., “Simultaneous harnessing of hot electrons and hot holes achieved via n-metal-p Janus plasmonic heteronanocrystals,” Nano Energy, vol. 98, p. 107217, 2022. https://doi.org/10.1016/j.nanoen.2022.107217.Search in Google Scholar
[50] G. Tagliabue, J. S. DuChene, A. Habib, R. Sundararaman, and H. A. Atwater, “Hot-hole versus hot-electron transport at Cu/GaN heterojunction interfaces,” ACS Nano, vol. 14, no. 5, pp. 5788–5797, 2020. https://doi.org/10.1021/acsnano.0c00713.Search in Google Scholar PubMed
[51] X. Li, H. O. Everitt, and J. Liu, “Synergy between thermal and nonthermal effects in plasmonic photocatalysis,” Nano Res., vol. 13, no. 5, pp. 1268–1280, 2020. https://doi.org/10.1007/s12274-020-2694-z.Search in Google Scholar
[52] X. Zhang, 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
[53] G. Baffou, I. Bordacchini, A. Baldi, and R. Quidant, “Simple experimental procedures to distinguish photothermal from hot-carrier processes in plasmonics,” Light: Sci. Appl., vol. 9, no. 1, p. 108, 2020. https://doi.org/10.1038/s41377-020-00345-0.Search in Google Scholar PubMed PubMed Central
[54] P. K. Jain, “Taking the heat off of plasmonic chemistry,” J. Phys. Chem. C, vol. 123, no. 40, pp. 24347–24351, 2019. https://doi.org/10.1021/acs.jpcc.9b08143.Search in Google Scholar
[55] E. Kazuma and Y. Kim, “Mechanistic studies of plasmon chemistry on metal catalysts,” Angew. Chem., Int. Ed., vol. 58, no. 15, pp. 4800–4808, 2019. https://doi.org/10.1002/anie.201811234.Search in Google Scholar PubMed
[56] M. Rodio, M. Graf, F. Schulz, N. S. Mueller, M. Eich, and H. Lange, “Experimental evidence for nonthermal contributions to plasmon-enhanced electrochemical oxidation reactions,” ACS Catal., vol. 10, no. 3, pp. 2345–2353, 2020. https://doi.org/10.1021/acscatal.9b05401.Search in Google Scholar
[57] T. E. Tesema, B. Kafle, M. G. Tadesse, and T. G. Habteyes, “Plasmon-enhanced resonant excitation and demethylation of methylene blue,” J. Phys. Chem. C, vol. 121, no. 13, pp. 7421–7428, 2017. https://doi.org/10.1021/acs.jpcc.7b00864.Search in Google Scholar
[58] Z. Fusco, A. Riaz, C. David, and F. J. Beck, “Cathodoluminescence spectroscopy of complex dendritic Au architectures for application in plasmon-mediated photocatalysis and as SERS substrates,” Adv. Mater. Interfaces, vol. 10, no. 3, p. 2202236, 2023. https://doi.org/10.1002/admi.202202236.Search in Google Scholar
[59] C. Brissaud, L. V. Besteiro, J. Y. Piquemal, and M. Comesaña-Hermo, “Plasmonics: a versatile toolbox for heterogeneous photocatalysis,” Sol. RRL, vol. 7, no. 13, p. 2300195, 2023. https://doi.org/10.1002/solr.202300195.Search in Google Scholar
[60] L. V. Besteiro, et al.., “Theory of plasmonic excitations,” in Plasmonic Catalysis, 2021, pp. 1–35.10.1002/9783527826971.ch1Search in Google Scholar
[61] Z. Zhu, R. Tang, C. Li, X. An, and L. He, “Promises of plasmonic antenna-reactor systems in gas-phase CO2 photocatalysis,” Adv. Sci., vol. 10, no. 24, p. 2302568, 2023. https://doi.org/10.1002/advs.202302568.Search in Google Scholar PubMed PubMed Central
[62] 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
[63] A. Y. Bykov, D. J. Roth, G. Sartorello, J. U. Salmón-Gamboa, and A. V. Zayats, “Dynamics of hot carriers in plasmonic heterostructures,” Nanophotonics, vol. 10, no. 11, pp. 2929–2938, 2021. https://doi.org/10.1515/nanoph-2021-0278.Search in Google Scholar
[64] L. Zhou, 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
[65] Y. Zhang, et al.., “Surface-plasmon-driven hot electron photochemistry,” Chem. Rev., vol. 118, no. 6, pp. 2927–2954, 2018. https://doi.org/10.1021/acs.chemrev.7b00430.Search in Google Scholar PubMed
[66] L. Yuan, B. B. Bourgeois, C. C. Carlin, F. H. da Jornada, and J. A. Dionne, “Sustainable chemistry with plasmonic photocatalysts,” Nanophotonics, vol. 12, no. 14, pp. 2745–2762, 2023. https://doi.org/10.1515/nanoph-2023-0149.Search in Google Scholar
[67] E. Cortes, et al.., “Plasmonic hot electron transport drives nano-localized chemistry,” Nat. Commun., vol. 8, p. 14880, 2017. https://doi.org/10.1038/ncomms14880.Search in Google Scholar PubMed PubMed Central
[68] D. Devasia, A. Das, V. Mohan, and P. K. Jain, “Control of chemical reaction pathways by light–matter coupling,” Annu. Rev. Phys. Chem., vol. 72, no. 1, pp. 423–443, 2021. https://doi.org/10.1146/annurev-physchem-090519-045502.Search in Google Scholar PubMed
[69] C. Boerigter, R. Campana, M. Morabito, and S. Linic, “Evidence and implications of direct charge excitation as the dominant mechanism in plasmon-mediated photocatalysis,” Nat. Commun., vol. 7, no. 1, p. 10545, 2016. https://doi.org/10.1038/ncomms10545.Search in Google Scholar PubMed PubMed Central
[70] J. J. Baumberg, “Hot electron science in plasmonics and catalysis: what we argue about,” Faraday Discuss., vol. 214, pp. 501–511, 2019. https://doi.org/10.1039/c9fd00027e.Search in Google Scholar PubMed
[71] H. Tang, et al.., “Plasmonic hot electrons for sensing, photodetection, and solar energy applications: a perspective,” J. Chem. Phys., vol. 152, no. 22, p. 220901, 2020. https://doi.org/10.1063/5.0005334.Search in Google Scholar PubMed
[72] 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
[73] Y. Sun and Z. Tang, “Photocatalytic hot-carrier chemistry,” MRS Bull., vol. 45, no. 1, pp. 20–25, 2020. https://doi.org/10.1557/mrs.2019.290.Search in Google Scholar
[74] S. Yu, A. J. Wilson, G. Kumari, X. Zhang, and P. K. Jain, “Opportunities and challenges of solar-energy-driven carbon dioxide to fuel conversion with plasmonic catalysts,” ACS Energy Lett., vol. 2, no. 9, pp. 2058–2070, 2017. https://doi.org/10.1021/acsenergylett.7b00640.Search in Google Scholar
[75] Y. Sivan, I. W. Un, and Y. Dubi, “Assistance of metal nanoparticles in photocatalysis – nothing more than a classical heat source,” Faraday Discuss., vol. 214, pp. 215–233, 2019. https://doi.org/10.1039/c8fd00147b.Search in Google Scholar PubMed
[76] Y. Dubi and Y. Sivan, ““Hot” electrons in metallic nanostructures-non-thermal carriers or heating?” Light: Sci. Appl., vol. 8, p. 89, 2019. https://doi.org/10.1038/s41377-019-0199-x.Search in Google Scholar PubMed PubMed Central
[77] F. B. Javier Aizpurua, et al.., “Theory of hot electrons: general discussion,” Faraday Discuss., vol. 214, p. 214, 2019.10.1039/C9FD90012HSearch in Google Scholar PubMed
[78] E.-R. Newmeyer, J. D. North, and D. F. Swearer, “Hot carrier photochemistry on metal nanoparticles,” J. Appl. Phys., vol. 132, no. 23, p. 230901, 2022. https://doi.org/10.1063/5.0123892.Search in Google Scholar
[79] Y. Dubi, I.-W. Un, and Y. Sivan, “Distinguishing Thermal from nonthermal (“Hot”) Carriers in illuminated molecular junctions,” Nano Lett., vol. 22, no. 5, pp. 2127–2133, 2022. https://doi.org/10.1021/acs.nanolett.1c04291.Search in Google Scholar PubMed
[80] Y. Dubi, I. W. Un, and Y. Sivan, “Thermal effects – an alternative mechanism for plasmon-assisted photocatalysis,” Chem. Sci., vol. 11, no. 19, pp. 5017–5027, 2020. https://doi.org/10.1039/c9sc06480j.Search in Google Scholar PubMed PubMed Central
[81] I. W. Un and Y. Sivan, “Parametric study of temperature distribution in plasmon-assisted photocatalysis,” Nanoscale, vol. 12, no. 34, pp. 17821–17832, 2020. https://doi.org/10.1039/d0nr03897k.Search in Google Scholar PubMed
[82] P. K. Jain, “Comment on “Thermal effects – an alternative mechanism for plasmon-assisted photocatalysis” by Y. Dubi, I. W. Un and Y. Sivan, Chem. Sci., 2020, 11, 5017,” Chem. Sci., vol. 11, no. 33, pp. 9022–9023, 2020. https://doi.org/10.1039/d0sc02914a.Search in Google Scholar PubMed PubMed Central
[83] X. Li, H. O. Everitt, and J. Liu, “Confirming nonthermal plasmonic effects enhance CO2 methanation on Rh/TiO2 catalysts,” Nano Res., vol. 12, no. 8, pp. 1906–1911, 2019. https://doi.org/10.1007/s12274-019-2457-x.Search in Google Scholar
[84] R. C. Elias and S. Linic, “Elucidating the roles of local and nonlocal rate enhancement mechanisms in plasmonic catalysis,” J. Am. Chem. Soc., vol. 144, no. 43, pp. 19990–19998, 2022. https://doi.org/10.1021/jacs.2c08561.Search in Google Scholar PubMed
[85] I.-W. Un and Y. Sivan, “The role of heat generation and fluid flow in plasmon-enhanced ReductioŽ oxidation reactions,” ACS Photonics, vol. 8, no. 4, pp. 1183–1190, 2021. https://doi.org/10.1021/acsphotonics.1c00113.Search in Google Scholar
[86] S. Mukherjee, 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
[87] P. Christopher, H. Xin, A. Marimuthu, and S. Linic, “Singular characteristics and unique chemical bond activation mechanisms of photocatalytic reactions on plasmonic nanostructures,” Nat. Mater., vol. 11, no. 12, pp. 1044–1050, 2012. https://doi.org/10.1038/nmat3454.Search in Google Scholar PubMed
[88] J. Huang, et al.., “Plasmon-induced optical control over dithionite-mediated chemical redox reactions,” Faraday Discuss., vol. 214, pp. 455–463, 2019. https://doi.org/10.1039/c8fd00155c.Search in Google Scholar PubMed PubMed Central
[89] 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
[90] A. A. Golubev, B. N. Khlebtsov, R. D. Rodriguez, Y. Chen, and D. R. T. Zahn, “Plasmonic heating plays a dominant role in the plasmon-induced photocatalytic reduction of 4-nitrobenzenethiol,” J. Phys. Chem. C, vol. 122, no. 10, pp. 5657–5663, 2018. https://doi.org/10.1021/acs.jpcc.7b12101.Search in Google Scholar
[91] T. E. Tesema, B. Kafle, and T. G. Habteyes, “Plasmon-driven reaction mechanisms: hot electron transfer versus plasmon-pumped adsorbate excitation,” J. Phys. Chem. C, vol. 123, no. 14, pp. 8469–8483, 2019. https://doi.org/10.1021/acs.jpcc.8b12054.Search in Google Scholar
[92] 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
[93] G. Baffou, et al.., “Photoinduced heating of nanoparticle arrays,” ACS Nano, vol. 7, no. 8, pp. 6478–6488, 2013. https://doi.org/10.1021/nn401924n.Search in Google Scholar PubMed
[94] C. Zhan, et al.., “Disentangling charge carrier from photothermal effects in plasmonic metal nanostructures,” Nat. Commun., vol. 10, no. 1, p. 2671, 2019. https://doi.org/10.1038/s41467-019-10771-3.Search in Google Scholar PubMed PubMed Central
[95] S. Tan, A. Argondizzo, J. Ren, L. Liu, J. Zhao, and H. Petek, “Plasmonic coupling at a metal/semiconductor interface,” Nat. Photonics, vol. 11, no. 12, pp. 806–812, 2017. https://doi.org/10.1038/s41566-017-0049-4.Search in Google Scholar
[96] E. Kazuma, J. Jung, H. Ueba, M. Trenary, and Y. Kim, “Real-space and real-time observation of a plasmon-induced chemical reaction of a single molecule,” Science, vol. 360, no. 6388, pp. 521–526, 2018. https://doi.org/10.1126/science.aao0872.Search in Google Scholar PubMed
[97] H. Robatjazi, 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
[98] E. Pensa, J. Gargiulo, A. Lauri, S. Schlücker, E. Cortés, and S. A. Maier, “Spectral screening of the energy of hot holes over a particle plasmon resonance,” Nano Lett., vol. 19, no. 3, pp. 1867–1874, 2019. https://doi.org/10.1021/acs.nanolett.8b04950.Search in Google Scholar PubMed
[99] E. Cortés, R. Grzeschik, S. A. Maier, and S. Schlücker, “Experimental characterization techniques for plasmon-assisted chemistry,” Nat. Rev. Chem, vol. 6, no. 4, pp. 259–274, 2022. https://doi.org/10.1038/s41570-022-00368-8.Search in Google Scholar PubMed
[100] D. Wang, et al.., “Spatial and temporal nanoscale plasmonic heating quantified by thermoreflectance,” Nano Lett., vol. 19, no. 6, pp. 3796–3803, 2019. https://doi.org/10.1021/acs.nanolett.9b00940.Search in Google Scholar PubMed
[101] D. F. Swearer, 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
[102] A. Carattino, M. Caldarola, and M. Orrit, “Gold nanoparticles as absolute nanothermometers,” Nano Lett., vol. 18, no. 2, pp. 874–880, 2018. https://doi.org/10.1021/acs.nanolett.7b04145.Search in Google Scholar PubMed PubMed Central
[103] M. Barella, et al.., “In situ photothermal response of single gold nanoparticles through hyperspectral imaging anti-Stokes thermometry,” ACS Nano, vol. 15, no. 2, pp. 2458–2467, 2021. https://doi.org/10.1021/acsnano.0c06185.Search in Google Scholar PubMed
[104] G. Baffou, “Anti-Stokes thermometry in nanoplasmonics,” ACS Nano, vol. 15, no. 4, pp. 5785–5792, 2021. https://doi.org/10.1021/acsnano.1c01112.Search in Google Scholar PubMed
[105] M. A.-V. Martin Holub, M. Adobes-Vidal, P. M. Gschwend, and D. Momotenko, “Andreas frutiger, pascal M. Gschwend, sotiris E. Pratsinis, and dmitry momotenko,” ACS Nano, vol. 14, no. 6, pp. 7358–7369, 2020. https://doi.org/10.1021/acsnano.0c02798.Search in Google Scholar PubMed
[106] B. Desiatov, I. Goykhman, and U. Levy, “Direct temperature mapping of nanoscale plasmonic devices,” Nano Lett., vol. 14, no. 2, pp. 648–652, 2014. https://doi.org/10.1021/nl403872d.Search in Google Scholar PubMed
[107] T. Shao, et al.., “A stacked plasmonic metamaterial with strong localized electric field enables highly efficient broadband light-driven CO2 hydrogenation,” Adv. Mater., vol. 34, no. 28, p. 2202367, 2022. https://doi.org/10.1002/adma.202202367.Search in Google Scholar PubMed
[108] E. Kazuma, J. Jung, H. Ueba, M. Trenary, and Y. Kim, “Direct pathway to molecular photodissociation on metal surfaces using visible light,” J. Am. Chem. Soc., vol. 139, no. 8, pp. 3115–3121, 2017. https://doi.org/10.1021/jacs.6b12680.Search in Google Scholar PubMed
[109] P. Christopher, H. Xin, and S. Linic, “Visible-light-enhanced catalytic oxidation reactions on plasmonic silver nanostructures,” Nat. Chem., vol. 3, no. 6, pp. 467–472, 2011. https://doi.org/10.1038/nchem.1032.Search in Google Scholar PubMed
[110] T. G. Habteyes, “Anions as intermediates in plasmon enhanced photocatalytic reactions,” J. Phys. Chem. C, vol. 124, no. 49, pp. 26554–26564, 2020. https://doi.org/10.1021/acs.jpcc.0c08831.Search in Google Scholar
[111] T. Habteyes, L. Velarde, and A. Sanov, “Photodissociation of CO2− in water clusters via Renner-Teller and conical interactions,” J. Chem. Phys., vol. 126, no. 15, p. 154301, 2007. https://doi.org/10.1063/1.2717932.Search in Google Scholar PubMed
[112] L. Nan, et al.., “Investigating plasmonic catalysis kinetics on hot-spot engineered nanoantennae,” Nano Lett., vol. 23, no. 7, pp. 2883–2889, 2023. https://doi.org/10.1021/acs.nanolett.3c00219.Search in Google Scholar PubMed
[113] M. Herran, et al.., “Tailoring plasmonic bimetallic nanocatalysts toward sunlight-driven H2 production,” Adv. Funct. Mater., vol. 32, no. 38, p. 2203418, 2022. https://doi.org/10.1002/adfm.202203418.Search in Google Scholar
[114] K. Li, N. J. Hogan, M. J. Kale, N. J. Halas, P. Nordlander, and P. Christopher, “Balancing near-field enhancement, absorption, and scattering for effective antenna–reactor plasmonic photocatalysis,” Nano Lett., vol. 17, no. 6, pp. 3710–3717, 2017. https://doi.org/10.1021/acs.nanolett.7b00992.Search in Google Scholar PubMed
[115] B. Seemala, 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
[116] R. Sundararaman, P. Narang, A. S. Jermyn, W. A. Goddard III, and H. A. Atwater, “Theoretical predictions for hot-carrier generation from surface plasmon decay,” Nat. Commun., vol. 5, p. 5788, 2014. https://doi.org/10.1038/ncomms6788.Search in Google Scholar PubMed PubMed Central
[117] A. M. Brown, R. Sundararaman, P. Narang, W. A. Goddard, 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
[118] A. Manjavacas, J. G. Liu, V. Kulkarni, and P. Nordlander, “Plasmon-induced hot carriers in metallic nanoparticles,” ACS Nano, vol. 8, no. 8, pp. 7630–7638, 2014. https://doi.org/10.1021/nn502445f.Search in Google Scholar PubMed
[119] 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
[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. Commun., vol. 6, no. 1, p. 7044, 2015. https://doi.org/10.1038/ncomms8044.Search in Google Scholar PubMed PubMed Central
[121] J. Fojt, T. P. Rossi, M. Kuisma, and P. Erhart, “Hot-carrier transfer across a nanoparticle–molecule junction: the importance of orbital hybridization and level alignment,” Nano Lett., vol. 22, no. 21, pp. 8786–8792, 2022. https://doi.org/10.1021/acs.nanolett.2c02327.Search in Google Scholar PubMed PubMed Central
[122] M. Kuisma, 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
[123] 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
[124] Z. Li, et al.., “Atomically defined undercoordinated copper active sites over nitrogen-doped carbon for aerobic oxidation of alcohols,” Small, vol. 18, no. 11, p. 2106614, 2022. https://doi.org/10.1002/smll.202106614.Search in Google Scholar PubMed
[125] C. Vogt and B. M. Weckhuysen, “The concept of active site in heterogeneous catalysis,” Nat. Rev. Chem, vol. 6, no. 2, pp. 89–111, 2022. https://doi.org/10.1038/s41570-021-00340-y.Search in Google Scholar PubMed
[126] P. V. Kumar, 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
[127] T. Le, Y. Shao, and B. Wang, “Plasmon-induced CO2 conversion on Al@Cu2O: a DFT study,” J. Phys. Chem. C, vol. 125, no. 11, pp. 6108–6115, 2021. https://doi.org/10.1021/acs.jpcc.0c10957.Search in Google Scholar
[128] Y. Zhang, et al.., “Indirect to direct charge transfer transition in plasmon-enabled CO2 photoreduction,” Adv. Sci., vol. 9, no. 2, p. 2102978, 2022. https://doi.org/10.1002/advs.202102978.Search in Google Scholar PubMed PubMed Central
[129] Y. Zhang, et al.., “Plasmon-mediated CO2 photoreduction via indirect charge transfer on small silver nanoclusters,” J. Phys. Chem. C, vol. 125, no. 48, pp. 26348–26353, 2021. https://doi.org/10.1021/acs.jpcc.1c07575.Search in Google Scholar
[130] A. J. Cohen, P. Mori-Sánchez, and W. Yang, “Challenges for density functional theory,” Chem. Rev., vol. 112, no. 1, pp. 289–320, 2012. https://doi.org/10.1021/cr200107z.Search in Google Scholar PubMed
[131] J. M. P. Martirez, J. L. Bao, and E. A. Carter, “First-principles insights into plasmon-induced catalysis,” Annu. Rev. Phys. Chem., vol. 72, no. 1, pp. 99–119, 2021. https://doi.org/10.1146/annurev-physchem-061020-053501.Search in Google Scholar PubMed
[132] K. P. Kuhl, T. Hatsukade, E. R. Cave, D. N. Abram, J. Kibsgaard, and T. F. Jaramillo, “Electrocatalytic conversion of carbon dioxide to methane and methanol on transition metal surfaces,” J. Am. Chem. Soc., vol. 136, no. 40, pp. 14107–14113, 2014. https://doi.org/10.1021/ja505791r.Search in Google Scholar PubMed
[133] A. A. Peterson and J. K. Nørskov, “Activity descriptors for CO2 electroreduction to methane on transition-metal catalysts,” J. Phys. Chem. Lett., vol. 3, no. 2, pp. 251–258, 2012. https://doi.org/10.1021/jz201461p.Search in Google Scholar
[134] Q. Zhao and E. A. Carter, “Revisiting competing paths in electrochemical CO2 reduction on copper via embedded correlated wavefunction theory,” J. Chem. Theory Comput., vol. 16, no. 10, pp. 6528–6538, 2020. https://doi.org/10.1021/acs.jctc.0c00583.Search in Google Scholar PubMed
[135] F. Libisch, C. Huang, and E. A. Carter, “Embedded correlated wavefunction schemes: theory and applications,” Acc. Chem. Res., vol. 47, no. 9, pp. 2768–2775, 2014. https://doi.org/10.1021/ar500086h.Search in Google Scholar PubMed
[136] J. L. Bao and E. A. Carter, “Rationalizing the hot-carrier-mediated reaction mechanisms and kinetics for ammonia decomposition on ruthenium-doped copper nanoparticles,” J. Am. Chem. Soc., vol. 141, no. 34, pp. 13320–13323, 2019. https://doi.org/10.1021/jacs.9b06804.Search in Google Scholar PubMed
[137] J. L. Bao and E. A. Carter, “Surface-plasmon-induced ammonia decomposition on copper: excited-state reaction pathways revealed by embedded correlated wavefunction theory,” ACS Nano, vol. 13, no. 9, pp. 9944–9957, 2019. https://doi.org/10.1021/acsnano.9b05030.Search in Google Scholar PubMed
[138] J. M. P. Martirez and E. A. Carter, “Excited-state N2 dissociation pathway on Fe-functionalized Au,” J. Am. Chem. Soc., vol. 139, no. 12, pp. 4390–4398, 2017. https://doi.org/10.1021/jacs.6b12301.Search in Google Scholar PubMed
[139] L. Zhou, 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
[140] Q. Zhao, J. M. P. Martirez, and E. A. Carter, “Revisiting understanding of electrochemical CO2 reduction on Cu(111): competing proton-coupled electron transfer reaction mechanisms revealed by embedded correlated wavefunction theory,” J. Am. Chem. Soc., vol. 143, no. 16, pp. 6152–6164, 2021. https://doi.org/10.1021/jacs.1c00880.Search in Google Scholar PubMed
[141] Z. Chen, Z. Liu, and X. Xu, “Accurate descriptions of molecule-surface interactions in electrocatalytic CO2 reduction on the copper surfaces,” Nat. Commun., vol. 14, no. 1, p. 936, 2023. https://doi.org/10.1038/s41467-023-36695-7.Search in Google Scholar PubMed PubMed Central
[142] Y. Wang, Y. Li, J. Chen, I. Y. Zhang, and X. Xu, “Doubly hybrid functionals close to chemical accuracy for both finite and extended systems: implementation and test of XYG3 and XYGJ-OS,” JACS Au, vol. 1, no. 5, pp. 543–549, 2021. https://doi.org/10.1021/jacsau.1c00011.Search in Google Scholar PubMed PubMed Central
[143] H. J. Huang, et al.., “Review of experimental setups for plasmonic photocatalytic reactions,” Catalysts, vol. 10, no. 5, pp. 1–25, 2020. https://doi.org/10.3390/catal10010046.Search in Google Scholar
[144] C. C. Carlin, et al.., “Nanoscale and ultrafast in situ techniques to probe plasmon photocatalysis,” Chem. Phys. Rev., vol. 4, no. 4, p. 041309, 2023. https://doi.org/10.1063/5.0163354.Search in Google Scholar
[145] E. Oksenberg, I. Shlesinger, G. Tek, A. F. Koenderink, and E. C. Garnett, “Complementary surface-enhanced Raman scattering (SERS) and IR absorption spectroscopy (SEIRAS) with nanorods-on-a-mirror,” Adv. Funct. Mater., vol. 33, no. 8, p. 2211154, 2023. https://doi.org/10.1002/adfm.202211154.Search in Google Scholar
[146] 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
[147] 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
[148] X. Chang, S. Vijay, Y. Zhao, N. J. Oliveira, K. Chan, and B. Xu, “Understanding the complementarities of surface-enhanced infrared and Raman spectroscopies in CO adsorption and electrochemical reduction,” Nat. Commun., vol. 13, no. 1, p. 2656, 2022. https://doi.org/10.1038/s41467-022-30262-2.Search in Google Scholar PubMed PubMed Central
[149] T. H. Tan, J. Scott, Y. H. Ng, R. A. Taylor, K. F. Aguey-Zinsou, and R. Amal, “C–C cleavage by Au/TiO2 during ethanol oxidation: understanding bandgap photoexcitation and plasmonically mediated charge transfer via quantitative in situ DRIFTS,” ACS Catal., vol. 6, no. 12, pp. 8021–8029, 2016. https://doi.org/10.1021/acscatal.6b01833.Search in Google Scholar
[150] E. R. Corson, et al.., “In situ ATR-SEIRAS of carbon dioxide reduction at a plasmonic silver cathode,” J. Am. Chem. Soc., vol. 142, no. 27, pp. 11750–11762, 2020. https://doi.org/10.1021/jacs.0c01953.Search in Google Scholar PubMed
[151] R. Kas, O. Ayemoba, N. J. Firet, J. Middelkoop, W. A. Smith, and A. Cuesta, “In-situ infrared spectroscopy applied to the study of the electrocatalytic reduction of CO2: theory, practice and challenges,” ChemPhysChem, vol. 20, no. 22, pp. 2904–2925, 2019. https://doi.org/10.1002/cphc.201900533.Search in Google Scholar PubMed
[152] X. Li, et al.., “Selective visible-light-driven photocatalytic CO2 reduction to CH4 mediated by atomically thin CuIn5S8 layers,” Nat. Energy, vol. 4, no. 8, pp. 690–699, 2019. https://doi.org/10.1038/s41560-019-0431-1.Search in Google Scholar
[153] E. Landaeta, N. I. Kadosh, and Z. D. Schultz, “Mechanistic study of plasmon-assisted in situ photoelectrochemical CO2 reduction to acetate with a Ag/Cu2O nanodendrite electrode,” ACS Catal., vol. 13, no. 3, pp. 1638–1648, 2023. https://doi.org/10.1021/acscatal.2c05082.Search in Google Scholar
[154] W. Shangguan, et al.., “Molecular-level insight into photocatalytic CO2 reduction with H2O over Au nanoparticles by interband transitions,” Nat. Commun., vol. 13, no. 1, p. 3894, 2022. https://doi.org/10.1038/s41467-022-31474-2.Search in Google Scholar PubMed PubMed Central
[155] H. Jia, Y. Dou, Y. Yang, F. Li, and C. y. Zhang, “Janus silver/ternary silver halide nanostructures as plasmonic photocatalysts boost the conversion of CO2 to acetaldehyde,” Nanoscale, vol. 13, no. 47, pp. 20289–20298, 2021. https://doi.org/10.1039/d1nr05801k.Search in Google Scholar PubMed
[156] C. Lu, et al.., “Constructing surface plasmon resonance on Bi2WO6 to boost high-selective CO2 reduction for methane,” ACS Nano, vol. 15, no. 2, pp. 3529–3539, 2021. https://doi.org/10.1021/acsnano.1c00452.Search in Google Scholar PubMed
[157] 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
[158] H. Song, et al.., “Light-enhanced carbon dioxide activation and conversion by effective plasmonic coupling effect of Pt and Au nanoparticles,” ACS Appl. Mater. Interfaces, vol. 10, no. 1, pp. 408–416, 2018. https://doi.org/10.1021/acsami.7b13043.Search in Google Scholar PubMed
[159] M. Sayed, L. Zhang, and J. Yu, “Plasmon-induced interfacial charge-transfer transition prompts enhanced CO2 photoreduction over Cu/Cu2O octahedrons,” Chem. Eng. J., vol. 397, no. 1, p. 125390, 2020. https://doi.org/10.1016/j.cej.2020.125390.Search in Google Scholar
[160] N. Ojha, A. K. Metya, and S. Kumar, “Influence of plasmonic metals (Ag, Cu) on overall CO2 photoreduction activity of β-Ga2O3,” Appl. Surf. Sci., vol. 580, p. 152315, 2022. https://doi.org/10.1016/j.apsusc.2021.152315.Search in Google Scholar
[161] Y. Cao, et al.., “Modulating electron density of vacancy site by single Au atom for effective CO2 photoreduction,” Nat. Commun., vol. 12, no. 1, p. 1675, 2021. https://doi.org/10.1038/s41467-021-21925-7.Search in Google Scholar PubMed PubMed Central
[162] L. Nguyen, F. F. Tao, Y. Tang, J. Dou, and X. J. Bao, “Understanding catalyst surfaces during catalysis through near ambient pressure X-ray photoelectron spectroscopy,” Chem. Rev., vol. 119, no. 12, pp. 6822–6905, 2019. https://doi.org/10.1021/acs.chemrev.8b00114.Search in Google Scholar PubMed
[163] L. Collado, et al.., “Unravelling the effect of charge dynamics at the plasmonic metal/semiconductor interface for CO2 photoreduction,” Nat. Commun., vol. 9, no. 1, p. 4986, 2018. https://doi.org/10.1038/s41467-018-07397-2.Search in Google Scholar PubMed PubMed Central
[164] C. Hu, et al.., “Near-infrared-featured broadband CO2 reduction with water to hydrocarbons by surface plasmon,” Nat. Commun., vol. 14, no. 1, p. 221, 2023. https://doi.org/10.1038/s41467-023-35860-2.Search in Google Scholar PubMed PubMed Central
[165] J. Zhao, et al.., “Plasmonic control of solar-driven CO2 conversion at the metal/ZnO interfaces,” Appl. Catal., B, vol. 256, no. 1, p. 117823, 2019. https://doi.org/10.1016/j.apcatb.2019.117823.Search in Google Scholar
[166] H. Liu, et al.., “Design of PdAu alloy plasmonic nanoparticles for improved catalytic performance in CO2 reduction with visible light irradiation,” Nano Energy, vol. 26, pp. 398–404, 2016. https://doi.org/10.1016/j.nanoen.2016.05.045.Search in Google Scholar
[167] C. Wang, et al.., “Endothermic reaction at room temperature enabled by deep-ultraviolet plasmons,” Nat. Mater., vol. 20, no. 3, pp. 346–352, 2021. https://doi.org/10.1038/s41563-020-00851-x.Search in Google Scholar PubMed PubMed Central
[168] 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
[169] K. Sytwu, 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
[170] F. Wang, et al.., “Plasmonic photocatalysis for CO2 reduction: advances, understanding and possibilities,” Chem. Eur. J., vol. 29, no. 25, 2023, Art. no. e202202716. https://doi.org/10.1002/chem.202202716.Search in Google Scholar PubMed
[171] R. Verma, R. Belgamwar, and V. Polshettiwar, “Plasmonic photocatalysis for CO2 conversion to chemicals and fuels,” ACS Mater. Lett., vol. 3, no. 5, pp. 574–598, 2021. https://doi.org/10.1021/acsmaterialslett.1c00081.Search in Google Scholar
[172] R.-G. Ciocarlan, N. Blommaerts, S. Lenaerts, P. Cool, and S. W. Verbruggen, “Recent trends in plasmon-assisted photocatalytic CO2 reduction,” ChemSusChem, vol. 16, no. 5, 2023, Art. no. e202201647. https://doi.org/10.1002/cssc.202201647.Search in Google Scholar PubMed
[173] S. Singh, et al.., “Surface plasmon-enhanced photo-driven CO2 hydrogenation by hydroxy-terminated nickel nitride nanosheets,” Nat. Commun., vol. 14, no. 1, p. 2551, 2023. https://doi.org/10.1038/s41467-023-38235-9.Search in Google Scholar PubMed PubMed Central
[174] X. Zhang, et al.., “MOF encapsulated sub-nm Pd skin/Au nanoparticles as antenna-reactor plasmonic catalyst for light driven CO2 hydrogenation,” Nano Energy, vol. 84, p. 105950, 2021. https://doi.org/10.1016/j.nanoen.2021.105950.Search in Google Scholar
[175] Q. Kang, et al.., “Photocatalytic reduction of carbon dioxide by hydrous hydrazine over Au–Cu alloy nanoparticles supported on SrTiO3/TiO2 coaxial nanotube arrays,” Angew. Chem., Int. Ed., vol. 54, no. 3, pp. 841–845, 2015. https://doi.org/10.1002/anie.201409183.Search in Google Scholar PubMed
[176] J. Becerra, D. T. Nguyen, V. N. Gopalakrishnan, and T. O. Do, “Plasmonic Au nanoparticles incorporated in the zeolitic imidazolate framework (ZIF-67) for the efficient sunlight-driven photoreduction of CO2,” ACS Appl. Energy Mater., vol. 3, no. 8, pp. 7659–7665, 2020. https://doi.org/10.1021/acsaem.0c01083.Search in Google Scholar
[177] M. Tahir, B. Tahir, and N. A. S. Amin, “Synergistic effect in plasmonic Au/Ag alloy NPs co-coated TiO2 NWs toward visible-light enhanced CO2 photoreduction to fuels,” Appl. Catal., B, vol. 204, pp. 548–560, 2017. https://doi.org/10.1016/j.apcatb.2016.11.062.Search in Google Scholar
[178] M. Tahir, B. Tahir, and N. A. S. Amin, “Gold-nanoparticle-modified TiO2 nanowires for plasmon-enhanced photocatalytic CO2 reduction with H2 under visible light irradiation,” Appl. Surf. Sci., vol. 356, pp. 1289–1299, 2015. https://doi.org/10.1016/j.apsusc.2015.08.231.Search in Google Scholar
[179] M. Tahir, “Synergistic effect in MMT-dispersed Au/TiO2 monolithic nanocatalyst for plasmon-absorption and metallic interband transitions dynamic CO2 photo-reduction to CO,” Appl. Catal., B, vol. 219, pp. 329–343, 2017. https://doi.org/10.1016/j.apcatb.2017.07.062.Search in Google Scholar
[180] M. Tahir, B. Tahir, N. A. S. Amin, and Z. Y. Zakaria, “Photo-induced reduction of CO2 to CO with hydrogen over plasmonic Ag-NPs/TiO2 NWs core/shell hetero-junction under UV and visible light,” J. CO2 Util., vol. 18, pp. 250–260, 2017. https://doi.org/10.1016/j.jcou.2017.02.002.Search in Google Scholar
[181] P. Martínez Molina, et al.., “Low temperature sunlight-powered reduction of CO2 to CO using a plasmonic Au/TiO2 nanocatalyst,” ChemCatChem, vol. 13, no. 21, pp. 4507–4513, 2021. https://doi.org/10.1002/cctc.202100699.Search in Google Scholar
[182] H. Robatjazi, 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
[183] 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
[184] E. Vahidzadeh, et al.., “Asymmetric multipole plasmon-mediated catalysis shifts the product selectivity of CO2 photoreduction toward C2+ products,” ACS Appl. Mater. Interfaces, vol. 13, no. 6, pp. 7248–7258, 2021. https://doi.org/10.1021/acsami.0c21067.Search in Google Scholar PubMed
[185] Q. Chen, et al.., “Photo-induced Au–Pd alloying at TiO2 {101} facets enables robust CO2 photocatalytic reduction into hydrocarbon fuels,” J. Mater. Chem. A, vol. 7, no. 3, pp. 1334–1340, 2019. https://doi.org/10.1039/c8ta09412h.Search in Google Scholar
[186] W. Tu, Y. Zhou, H. Li, P. Li, and Z. Zou, “Au@TiO2 yolk–shell hollow spheres for plasmon-induced photocatalytic reduction of CO2 to solar fuel via a local electromagnetic field,” Nanoscale, vol. 7, no. 34, pp. 14232–14236, 2015. https://doi.org/10.1039/c5nr02943k.Search in Google Scholar PubMed
[187] M. Bonchio, et al.., “Best practices for experiments and reporting in photocatalytic CO2 reduction,” Nat. Catal., vol. 6, no. 8, pp. 657–665, 2023. https://doi.org/10.1038/s41929-023-00992-7.Search in Google Scholar
[188] A. Olivo, E. Ghedini, M. Signoretto, M. Compagnoni, and I. Rossetti, “Liquid vs. Gas phase CO2 photoreduction process: which is the effect of the reaction medium?” Energies, vol. 10, no. 9, pp. 1–14, 2017. https://doi.org/10.3390/en10091394.Search in Google Scholar
[189] C.-Y. Chen, J. Yu, V. H. Nguyen, J. Wu, W. H. Wang, and K. Kočí, “Reactor design for CO2 photo-hydrogenation toward solar fuels under ambient temperature and pressure,” Catalysts, vol. 7, no. 12, pp. 1–12, 2017. https://doi.org/10.3390/catal7020063.Search in Google Scholar
[190] S. Ali, et al.., “Gas phase photocatalytic CO2 reduction, “A brief Overview for benchmarking”,” Catalysts, vol. 9, no. 9, pp. 1–26, 2019. https://doi.org/10.3390/catal9090727.Search in Google Scholar
[191] A. Álvarez, et al.., “CO2 activation over catalytic surfaces,” ChemPhysChem, vol. 18, no. 22, pp. 3135–3141, 2017. https://doi.org/10.1002/cphc.201700782.Search in Google Scholar PubMed
[192] C. Kim, et al.., “Energy-efficient CO2 hydrogenation with fast response using photoexcitation of CO2 adsorbed on metal catalysts,” Nat. Commun., vol. 9, no. 1, p. 3027, 2018. https://doi.org/10.1038/s41467-018-05542-5.Search in Google Scholar PubMed PubMed Central
[193] L. Meng, et al.., “Gold plasmon-induced photocatalytic dehydrogenative coupling of methane to ethane on polar oxide surfaces,” Energy Environ. Sci., vol. 11, no. 2, pp. 294–298, 2018. https://doi.org/10.1039/c7ee02951a.Search in Google Scholar
[194] Y. Liu, Q. Chen, D. A. Cullen, Z. Xie, and T. Lian, “Efficient hot electron transfer from small Au nanoparticles,” Nano Lett., vol. 20, no. 6, pp. 4322–4329, 2020. https://doi.org/10.1021/acs.nanolett.0c01050.Search in Google Scholar PubMed
[195] D. Esrafilzadeh, et al.., “Room temperature CO2 reduction to solid carbon species on liquid metals featuring atomically thin ceria interfaces,” Nat. Commun., vol. 10, no. 1, p. 865, 2019. https://doi.org/10.1038/s41467-019-08824-8.Search in Google Scholar PubMed PubMed Central
[196] P. Gao, et al.., “Direct conversion of CO2 into liquid fuels with high selectivity over a bifunctional catalyst,” Nat. Chem., vol. 9, no. 10, pp. 1019–1024, 2017. https://doi.org/10.1038/nchem.2794.Search in Google Scholar PubMed
[197] N. Han, P. Liu, J. Jiang, L. Ai, Z. Shao, and S. Liu, “Recent advances in nanostructured metal nitrides for water splitting,” J. Mater. Chem. A, vol. 6, no. 41, pp. 19912–19933, 2018. https://doi.org/10.1039/c8ta06529b.Search in Google Scholar
[198] F. Song, W. Li, J. Yang, G. Han, P. Liao, and Y. Sun, “Interfacing nickel nitride and nickel boosts both electrocatalytic hydrogen evolution and oxidation reactions,” Nat. Commun., vol. 9, no. 1, p. 4531, 2018. https://doi.org/10.1038/s41467-018-06728-7.Search in Google Scholar PubMed PubMed Central
[199] H. Dai, H. He, P. Li, L. Gao, and C. T. Au, “The relationship of structural defect–redox property–catalytic performance of perovskites and their related compounds for CO and NOx removal,” Catal. Today, vol. 90, no. 3, pp. 231–244, 2004. https://doi.org/10.1016/j.cattod.2004.04.031.Search in Google Scholar
[200] U. Ulmer, et al.., “Fundamentals and applications of photocatalytic CO2 methanation,” Nat. Commun., vol. 10, no. 1, p. 3169, 2019. https://doi.org/10.1038/s41467-019-10996-2.Search in Google Scholar PubMed PubMed Central
[201] K. Hirai, J. A. Hutchison, and H. Uji-i, “Recent progress in vibropolaritonic chemistry,” ChemPlusChem, vol. 85, no. 9, pp. 1981–1988, 2020. https://doi.org/10.1002/cplu.202000411.Search in Google Scholar PubMed
[202] R. Chikkaraddy, et al.., “Single-molecule strong coupling at room temperature in plasmonic nanocavities,” Nature, vol. 535, no. 7610, pp. 127–130, 2016. https://doi.org/10.1038/nature17974.Search in Google Scholar PubMed PubMed Central
[203] S. Ma and Z.-P. Liu, “Machine learning for atomic simulation and activity prediction in heterogeneous catalysis: current status and future,” ACS Catal., vol. 10, no. 22, pp. 13213–13226, 2020. https://doi.org/10.1021/acscatal.0c03472.Search in Google Scholar
[204] D. Chen, C. Shang, and Z.-P. Liu, “Machine-learning atomic simulation for heterogeneous catalysis,” npj Comput. Mater., vol. 9, no. 1, p. 2, 2023. https://doi.org/10.1038/s41524-022-00959-5.Search in Google Scholar
© 2024 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
- Reviews
- Advances in fundamentals and application of plasmon-assisted CO2 photoreduction
- Optical computing metasurfaces: applications and advances
- Research Articles
- Highly efficient upconversion photodynamic performance of rare-earth-coupled dual-photosensitizers: ultrafast experiments and excited-state calculations
- Imaging the scattered light of a nanoparticle through a cylindrical capillary
- Dual high-Q Fano resonances metasurfaces excited by asymmetric dielectric rods for refractive index sensing
- Multi-wavelength structured light based on metasurfaces for 3D imaging
- Interactions and ultrafast dynamics of exciton complexes in a monolayer semiconductor with electron gas
- Full-polarization-locked vortex beam generator with time-varying characteristics
- Topological edge and corner states in coupled wave lattices in nonlinear polariton condensates
- Spatial signature of the photoelastic effect in the acoustic–plasmonic coupling revealed by space responsivity induced by polarized optical excitation
- Cylindrical vector beam multiplexing holography employing spin-decoupled phase modulation metasurface
Articles in the same Issue
- Frontmatter
- Reviews
- Advances in fundamentals and application of plasmon-assisted CO2 photoreduction
- Optical computing metasurfaces: applications and advances
- Research Articles
- Highly efficient upconversion photodynamic performance of rare-earth-coupled dual-photosensitizers: ultrafast experiments and excited-state calculations
- Imaging the scattered light of a nanoparticle through a cylindrical capillary
- Dual high-Q Fano resonances metasurfaces excited by asymmetric dielectric rods for refractive index sensing
- Multi-wavelength structured light based on metasurfaces for 3D imaging
- Interactions and ultrafast dynamics of exciton complexes in a monolayer semiconductor with electron gas
- Full-polarization-locked vortex beam generator with time-varying characteristics
- Topological edge and corner states in coupled wave lattices in nonlinear polariton condensates
- Spatial signature of the photoelastic effect in the acoustic–plasmonic coupling revealed by space responsivity induced by polarized optical excitation
- Cylindrical vector beam multiplexing holography employing spin-decoupled phase modulation metasurface