Open Access Article
Wen-Qian Lia,
Miao Xub,
Gang Chencd,
Xiaoping Chen
*cd,
Jie-Sheng Chen
a and
Tian-Nan Ye
*a
aFrontiers Science Center for Transformative Molecules, School of Chemistry and Chemical Engineering, Shanghai Jiao Tong University, Shanghai 200240, China. E-mail: ytn2011@sjtu.edu.cn
bState Key Laboratory of Space Power-Sources Technology, Shanghai Institute of Space Power-Sources, Shanghai 200245, China
cInstitute of Energy Research, Jiangxi Academy of Sciences, Nanchang 330096, China. E-mail: cxpjxskxy@126.com
dJiangxi Carbon Neutralization Research Centre, Jiangxi Academy of Sciences, Nanchang 330096, China
First published on 13th December 2025
Ammonia is an important component in the manufacture of fertilizers and various chemicals, and its production mainly relies on the energy-intensive Haber–Bosch process. To overcome this, there has been growing interest in using photocatalysis as an alternative approach for ammonia synthesis under ambient conditions. Plasmonic nanomaterials have been considered to be particularly promising due to their localized surface plasmon resonance (LSPR) effects that combine the advantages of photochemical and thermal properties in one system. This review introduces the fundamental principles of LSPR effects, including hot carrier injection, photoheating and near-field enhancement. It then undertakes a comprehensive analysis of the current state-of-the-art catalysts for plasmon-driven photocatalytic ammonia synthesis. Finally, it proposes a brief outlook on the strategies for the design of plasmonic photocatalysts, advances in in situ characterization and theoretical simulations, standardization of the reaction conditions and detection technologies for ammonia production.
The pursuit of such sustainable strategies has catalysed the emergence of the “green ammonia” concept, which aims to fully decarbonize the ammonia synthesis process by leveraging renewable energy sources.7,8 A key step towards this goal is to replace natural gas-derived H2 with water as the hydrogen source to produce NH3 (2N2 + 6H2O → 4NH3 + 3O2).9,10 Photo-catalysis offers a promising solution for this purpose, as it enables ammonia synthesis at lower temperatures and pressures while utilizing renewable energy sources. In the early 1970s, Schrauzer and Guth reported the first instance of photo-catalytic ammonia synthesis. Since then, extensive research has been devoted to developing efficient catalysts for photocatalytic ammonia synthesis (Fig. 1).11 During the photocatalysis process, exposure to solar light causes the absorption of light energy, leading to a separation of the electron–hole pairs. Electrons are excited into the conduction band (CB) of the semiconductor for N2 reduction. Meanwhile, holes in the valence band (VB) facilitate the oxidation of H2O to O2 as well as the release of protons that combine with N atoms to form NH3. To date, various semiconductor materials have been explored as photocatalysts for ammonia synthesis, including oxides,12 sulfides,13 bismuth oxyhalides,9 carbonaceous materials14 and so on. Recently, it was reported that defective pyrochlore K2Ta2O6−x, loaded with Ru and characterized by a very low work function, facilitated efficient photocatalytic ammonia synthesis.12 The effectiveness of this process is primarily due to a synergistic effect: efficient charge separation, paired with spontaneous electron transfer, which is enhanced by the presence of oxygen defects. In addition to metal oxides, metal sulfides have attracted attention owing to their ability to absorb visible light.15 CdS was the first metal sulfide applied in photocatalytic ammonia synthesis, followed by other metal sulfides such as FeMoS chalcogels and Fe-decorated 2D-MoS2.13,16 However, the low physicochemical stability of metal sulfides hinders their further development. In recent years, two-dimensional materials such as bismuth oxyhalides,9 g-C3N4
14 and LDHs17 have gained attention in photocatalytic ammonia synthesis due to their quick electron transfer and large specific surface area. Bismuth oxyhalides have an internal electric field, which can promote carrier separation, while g-C3N4 was the first organic material used for photocatalytic ammonia synthesis, opening up research on organic photocatalysts. LDHs have a facilely controllable metal cation composition, allowing for the easy construction of defect and regulation of the band structure. Despite such encouraging progress, most photocatalysts continue to face challenges in achieving efficient photocatalytic N2 reduction (Fig. 2): (1) inherent limitations of semiconductor photocatalysts in light harvesting and utilization; (2) the undesirable recombination of excited electron–hole pairs that decreases the catalytic efficiency; (3) inadequate active sites and sluggish reaction kinetics that hamper N2 activation. Consequently, there is still a pressing need to broaden the scope of light-driven catalysis and enhance the efficiency of photocatalysts.
![]() | ||
| Fig. 2 The history, innovations, and challenges of ammonia synthesis through thermochemical, photochemical, and photo-thermal processes. | ||
Plasmonic photocatalysis has emerged as a promising technique for driving various chemical reactions, including CO2 reduction, water-splitting reactions, and organic compound transformations.18–24 Compared to semiconductors, plasmonic nanomaterials can absorb light across a wider range of the solar spectrum, from ultraviolet spectrum (UV) to near-infrared spectrum (NIR). The localized surface plasmon resonance (LSPR) of plasmonic nanomaterials, such as Ag, Au, and Cu nanoparticles, contributes to the high efficiency of plasmonic photocatalysis.25 On the one hand, the photo-induced electrons originating from plasmonic materials are transferred to the antibonding π orbital of N2 to activate N2. On the other hand, the photoheating effect can reduce the energy barrier for N2 activation. In addition, the enhanced electric field derived from plasmonic materials can polarize the N
N bond, facilitating the N2 dissociation (Fig. 2).26 By utilizing solar energy, this photo-thermal catalysis combines photochemical and thermochemical processes, leading to exceptional catalytic activity under mild conditions. To date, only a few reviews have reported on plasmon-enhanced ammonia synthesis. However, the focus of those reviews was on the state-of-the-art plasmonic catalysts rather than on the promotion mechanism for the ammonia synthesis process.27–29 In this review, we aim to provide a systematic summary of the fundamental principles of LSPR and to clearly demonstrate the promotion effects of LSPR in plasmonic-based ammonia synthesis. Herein, we will present the basics of LSPR and its underlying mechanisms, followed by a detailed examination of strategies for ammonia synthesis via plasmonic photocatalysis. Ultimately, we will conclude with a brief outlook on future directions.
However, charge carriers generated from plasmon decay have very short lifetimes of 1–100 fs, much shorter than those in semiconductors and molecular photocatalysts.33 To utilize these short-lived carriers in catalysis, they must be either extracted or leveraged.34 Two mechanisms for charge transfer have been identified: indirect charge transfer, where energetic carriers are generated in the plasmonic metal and later transferred to an adsorbate's unoccupied molecular orbital or a semiconductor's conduction band, and direct charge transfer, which involves the immediate excitation of charge carriers from the metal to hybridized metal–adsorbate interfacial states during plasmon decay.
To harness the full potential of hot carriers generated by plasmonic materials, several approaches have been employed in the design of plasmon-based catalysts. The first strategy involves constructing a hybrid structure of plasmon and semiconductor particles. In this design, plasmonic nanoparticles act as electron traps, transferring hot electrons to the conduction band of the semiconductor through a Schottky barrier (Fig. 4a).35 However, only a limited number of energetic electrons can be successfully transferred due to the high Schottky barrier (A Schottky barrier is a metal–semiconductor junction that exhibits rectifying behaviour, meaning it allows current to flow more easily in one direction than the other.) (Fig. 4b). To overcome this barrier, it is essential to establish a suitable interface between the plasmonic nanoparticles and the semiconductors. The second strategy involves combining two different metals (Fig. 4c). In this approach, plasmonic metal generates hot electrons that are transferred to the active transition metal, which provides active sites for N2 adsorption and activation. This approach is closer to an ohmic contact (an ohmic contact is a metal–semiconductor junction that exhibits linear or Ohm's law behaviour, meaning it allows current flow equally well in both directions), which requires a lower energy barrier for electron transfer.35
![]() | ||
| Fig. 4 (a) The scheme of photocatalytic mechanism over a metal–semiconductor system. (b) The scheme of Schottky contact. (c) The scheme of ohmic contact. (d) Calculated temperature increase in the centre of a square array of 16 NPs. (e) Calculated temperature increase on the surface of a single Au NP as a function of illumination power at the plasmon resonance. Reproduced with permission43. Copyright 2007, Elsevier. (f) Electric field enhancement of the Au nano-bipyramid and Au nanorod. Reproduced with permission44. Copyright 2007, Elsevier. | ||
The photoheating effect is minimal when applied to a single metal nanoparticle.45 However, as the number of nanoparticles (NPs) increased, the photoheating effect can significantly enhance due to the accumulative effect.43 Fig. 4d shows the calculated temperature increase in the centre of a square array of 16 NPs. It is evident that the more the number of NPs, the stronger the temperature increase observed in the system. The photoheating effect is also highly dependent on the particle size of plasmonic metals. As shown in Fig. 4e, the NP with a larger radius shows a stronger temperature increase.43 Photoheating is a ubiquitous phenomenon in plasmonic systems, significantly enhancing the dynamic process and increasing the yield of the product. The use of temperature-sensitive photocatalysts offers promising opportunities to harness the photoheating effect.
![]() | ||
Fig. 5 (a) Power-dependent study of CuPtSA (246 : 1) disentangling photothermal and nonthermal effects on plasmonic propane dehydrogenation at 460 °C. Reproduced with permission46. Copyright 2025, American Chemical Society. (b) Absorption and AQE action spectra of the Au NBP/Rh SSs. Reproduced with permission47. Copyright 2024, American Chemical Society. (c) The corresponding fs-TAS decay curves at 500 nm for various samples in the lactic acid solution. Reproduced with permission48. Copyright 2025, American Chemical Society. (d) Energy transfer pathways for photomediated chemistry on a plasmonic metal nanoparticle catalyst. Reproduced with permission49. Copyright 2024, American Chemical Society. (e) Raman frequency dependence for Nb–O stretching vibrations. Reproduced with permission50. Copyright 2017, Wiley-VCH. (f) Cs–Ru/MgO for photocatalytic ammonia synthesis. Reproduced with permission51. Copyright 2019, The American Chemical Society. (g) FDTD analysis of Au NBPs. Reproduced with permission47. Copyright 2024, American Chemical Society. (h) FDTD simulation of electric field intensity distribution in the Pt–Au/SiO2 system. Reproduced with permission52. Copyright 2018, American Chemical Society. | ||
The photocatalytic rate's dependence on light intensity is a powerful analytical tool for understanding the reaction mechanisms. Four kinetic categories related to the light intensity (I) and photocatalytic reaction rate have been documented: sublinear (rates ∝ In, n <1), linear (rates ∝ I), superlinear (rates ∝ In, n >1), and exponential (rates ∝ ef(I)). Sublinear dependence is typical in non-plasmonic semiconductors where charge carrier recombination is dominant. A linear reaction rate with light intensity is a hallmark of electron-driven processes, commonly reported in plasmonic reactions (Fig. 5a).46 Superlinear behavior may also suggest hot carrier-driven transformations, where multiple excitations of vibrational modes by hot electrons lead to enhanced reactions. Conversely, if photothermal effects dominate, the reaction shows Arrhenius-type behavior, exhibiting exponential dependence on light intensity.
The dependence of reaction rates on photon wavelength can further indicate hot carrier-driven processes (Fig. 5b). A correlation between the absorption spectrum of the plasmonic photocatalyst and wavelength-dependent reaction rates supports electron-driven photochemical transformations.47 Transient absorption spectroscopy (TAS) is effective for studying electron transfer dynamics (Fig. 5c).48 Rapid electron transfers occur within hundreds of femtoseconds, while the transient signal's decay lasts thousands of nanoseconds, indicating effective charge carrier separation and reduced recombination. Other techniques, like photoluminescence analysis and photovoltage measurements, can also identify hot electron transfer and charge separation.
For photothermal effects, direct temperature measurements at active sites are critical (Fig. 5d).49 Recent advancements in nanoscale thermometry include non-luminescence methods, like scanning thermal microscopy, which achieves a spatial resolution of 10 nm with 10–50 mK precision, and luminescence methods such as thermoreflectance and optical interferometry. Thermoreflectance utilizes the correlation between a material's refractive index and temperature, creating temperature profiles with high resolution. Additionally, tip-enhanced Raman spectroscopy (TERS) can map local temperatures at the nanometer scale by calculating the ratio of anti-Stokes to Stokes Raman signal intensities. For example, Ozin et al. investigated a Pd/Nb2O5 catalyst for CO2 photocatalytic reduction, using Stokes and anti-Stokes Raman bands to determine the local catalyst temperature (Fig. 5e).50 Their findings suggested that Pd NPs act as photothermal “nano-heaters”, effectively raising local temperatures to enhance CO2 hydrogenation. Controlled experiments are also crucial for identifying photothermal effects. For example, Li et al. demonstrated that using a black photothermal material could completely suppress hot electron transfer, converting all light into heat (Fig. 5f).51
To identify local near-field-driven plasmonic photocatalysis mechanisms, various methods have been employed. The finite difference time domain (FDTD) method is a widely used numerical tool for full-wave electromagnetic field analysis. In plasmonic photocatalysis, FDTD simulations can evaluate the interaction between light and nanoparticles, providing insights into optimizing systems through near-field distribution and intensity enhancement.46 Recently, Yang et al. used FDTD to simulate electric field enhancement from Au nano-bipyramids (Fig. 5g), analyzing physical parameters such as the aspect ratio and excitation wavelength.47 Additionally, constructing bimetallic plasmonic photocatalysts can enhance plasmonic coupling between materials, leading to significant electric fields and efficient electron transfer (Fig. 5h).52
A highly effective strategy involves the introduction of defects, particularly vacancies, in semiconductor supports or co-catalysts. These vacancies function as effective electron traps, capturing plasmon-derived hot electrons and thereby mitigating charge recombination. Concurrently, they often serve as active sites for molecular adsorption and activation.9,56,57 For instance, oxygen vacancies in TiO2 have been extensively studied in this field. A representative system involves Au nanoparticles anchored on TiO2 nanosheets with sufficient oxygen vacancies, which were synthesized by the solvothermal method.53 In this architecture, the Au NPs act as optical antennas, generating hot electrons upon irradiation, which are subsequently transferred to the TiO2 conduction band and ultimately captured by the oxygen vacancies. These vacancy-trapped electrons then drive the activation and reduction of adsorbed N2 molecules (Fig. 6a). However, the oxygen vacancies are generated by the solvothermal method, a process that creates defects in the bulk phase. These defects may act as carrier traps, leading to the recombination of photogenerated electron–hole pairs, significantly reducing the overall quantum efficiency. Simultaneously, bulk defects are located within the material's interior, limiting access to reactant molecules and thereby reducing the density of active sites. In contrast, surface defects circumvent these limitations, effectively trapping photogenerated electrons to suppress bulk recombination while enabling rapid transfer to surface adsorbates. This direct pathway enhances interfacial charge transfer and improves surface reaction kinetics. This principle was demonstrated in a system where surface oxygen vacancies were precisely created on plasmonic TiO2/Au nanorods via atomic layer deposition.54 The confinement of vacancies to the surface region preserved the bulk properties of TiO2 and provided abundant, accessible active sites for N2 activation, leading to a significantly enhanced NH3 yield compared to the system with bulk defects (Fig. 6b).
![]() | ||
| Fig. 6 (a) Schematic illustration of the hot-electron generation, injection and N2 reduction over Au/TiO2 with oxygen vacancies. Reproduced with permission.53 Copyright 2018, American Chemical Society. (b) An illustration of the synergistic effect of surface oxygen vacancies and plasmonic Au NPs for ammonia synthesis. Reproduced with permission.54 Copyright 2018, Wiley-VCH. (c) Schematic illustration of the plasmonic photocatalytic ammonia synthesis. (d) NH3 yield on Au/TiO2 with different alkali metal cations. (e) Cathodic photocurrent on Au/TiO2 with different alkali metal cations. Reproduced with permission.55 Copyright 2019, The Royal Society of Chemistry. | ||
Beyond optimizing material-based electron acceptors, the reaction microenvironment at the solid–liquid interface can be strategically engineered to enhance plasmon-mediated N2 reduction further. Introducing alkali and alkaline earth metal (AM) cations (e.g., K+, Na+, etc.) into the reaction system has proven to significantly boost catalytic activity (Fig. 6c–e), building on the foundation provided by optimized electron acceptors like oxygen vacancies.55 The promotional effect of these cations is primarily due to the formation of a strong local electric field at the catalyst–solution interface, rather than direct involvement in the charge trapping process. It is proposed that this cation-induced electric field effectively stabilizes key intermediates generated during the N2 reduction process, subsequently lowering the activation energy barrier for the rate-determining step.49 However, this mechanism remains speculative, as direct evidence for the presence and role of such a field is lacking. In a related study on electrocatalytic CO2 reduction, Chen et al. developed a microkinetic model incorporating electric field effects from ab initio calculations. Their work demonstrated that the field from solvated cations in the double layer, along with their image charges on the metal surface, can significantly stabilize critical intermediates like *CO2.58 These findings support the notion that cations can indeed influence key intermediates through local field effects. To gain deeper insights into the cation-induced electric field mechanism, future investigations will require more experimental evidence, such as advanced in situ FT-IR and in situ XPS measurements capable of probing the interaction between the electric field and reaction intermediates in real-time.
The strategy of optimizing electron acceptors can be extended beyond TiO2 to other reducible oxides, such as ceria (CeO2), which features a reversible Ce4+/Ce3+ redox couple associated with oxygen vacancy formation. A noteworthy example is the site-selective growth of Au nanorods capped with crystalline CeO2 at their ends.50 This unique spatial configuration, with oxygen vacancies localized in the ceria domains, facilitates improved charge separation. Plasmon-induced hot electrons generated in the Au nanorods are efficiently transferred to the oxygen vacancy sites of CeO2 for N2 activation, while the hot holes are scavenged by a sacrificial agent. As a result, this system demonstrates enhanced N2 fixation activity under near-infrared illumination (Fig. 7a–c). Another interesting work reported that abundant surface vacancies of CeO2 can effectively stabilize the individual Ru atoms.60 Under illumination, CeO2 generates electron–hole pairs, with electron captures by the Ru sites. The electron-rich Ru site can effectively activate N2 molecules, which reduce the energy barrier of the rate-limiting step, thereby enhancing the ammonia synthesis performance.
![]() | ||
| Fig. 7 (a) Schematic illustration of the synthesis of Au/end CeO2. (b) Mechanism of N2 photofixation for Au/end-CeO2. (c) Comparison of the hot carrier separation behaviours of Au/end-CeO2 with that of the core@shell nanostructure. Reproduced with permission.59 Copyright 2019, American Chemical Society. | ||
The scope of plasmonic materials has expanded to include certain inorganic compounds beyond metal nanoparticles, such as Ti3C2Tx, MoO3, and SrMoO4. Ti3C2Tx is a typical 2D MXene material,61–64 which shows intrinsic LSPR in the visible to near-infrared range.65,66 This LSPR behavior fundamentally arises from the material's unique electronic structure, which is directly reflected in its optical constants (ñ = n + ik). The high concentration of Drude-like free electrons inherent to the metallic Ti3C2 core results in a large negative real part of the dielectric function (ε1) in the visible-NIR region,67 which is essential for LSPR excitation. Additionally, significant optical absorption, characterized by a substantial extinction coefficient (k), leads to the observed strong LSPR peaks. The surface functional groups (Tx = –O, –OH, and –F) critically modulate these optical constants by acting as n-type dopants. By withdrawing the electron density from the titanium layers, these terminations effectively increase the free carrier concentration. This, in turn, alters the plasma frequency and directly tunes the values of n and k, thereby enabling precise engineering of the LSPR wavelength and intensity. Consequently, the ability to tailor the optical constants through composition and structure makes Ti3C2Tx a highly versatile and rationally designable platform for plasmon-enhanced applications. Similar to the LSPR metals, single Ti3C2Tx often shows inadequate charge carrier separation efficiency, limiting its usefulness in plasmonic photocatalysis. A promising approach to solve this dilemma involves creating vacancies to hinder carrier recombination. Hou et al. designed Ti3C2Tx/TiO2 hybrid structures, in which Ti3C2Tx serves as the plasmonic component, triggering the generation of hot electrons upon light absorption, while the oxygen vacancies in TiO2 act as catalytically active sites for N2 activation.68 Such a hybrid structure (Fig. 8a) greatly enhances the efficiency of photocarriers, leading to excellent performance in photocatalytic ammonia synthesis.
![]() | ||
| Fig. 8 (a) Schematic illustration of the electronic structures for Ti3C2Tx/TiO2-400. Reproduced with permission.68 Copyright 2020, Elsevier. (b) Schematic illustration of plasmonic photocatalytic N2 fixation over MoO3−x. Reproduced with permission.69 Copyright 2020, The Royal Society of Chemistry. (c) The structure of SrMoO4 with oxygen vacancies. (d) The mechanism for plasmonic photocatalytic nitrogen fixation over SrMoO4 with oxygen vacancies. Reproduced with permission.70 Copyright 2021, Elsevier. | ||
In addition to MXenes, some non-stoichiometric oxides, such as MoO3−x and SrMoO4, also exhibit LSPR behaviour. For MoO3−x, the LSPR is primarily driven by the introduction of oxygen vacancies. These vacancies act as intrinsic donors, generating a high concentration of free electrons by reducing Mo6+ to Mo5+ states. The collective oscillation of these vacancy-induced free carriers upon photoexcitation results in a strong, tunable LSPR absorption. Recently, a single MoO3−x nanosheet with a high charge carrier density of 7.46 × 1020 cm−3 was reported to achieve the N2 photofixation in pure water under visible to NIR plasmonic excitation. Abundant OVs and Mo5+ centres coexist in this “two-in-one” semiconductor, enabling the introduction of rich active sites and broad spectrum LSPR-induced hot electrons with a high reductive potential into one nanostructure (Fig. 8b).69 Such a strategy endows the MoO3−x nanosheet a high ammonia synthesis rate without the need for any other co-catalyst, providing a new route for the design and fabrication of plasmonic semiconductors.
Similarly, plasmonic SrMoO4 can also be obtained by annealing SrMoO4 in hydrogen, creating abundant oxygen vacancies with a high charge carrier density of ∼2.0 × 1020 cm−3 and leading to add-on states close to the Fermi level (Fig. 8c).70 Upon excitation, plasmonic hot electrons are transformed to the conduction band (Fig. 8d), allowing efficient harvesting of the visible and near-infrared light. These hot electrons are energetic enough to trigger the photocatalytic nitrogen reduction reaction. As a result, the plasmonic SrMoO4 shows enhanced photocatalytic nitrogen reduction performance under irradiation of visible and near-infrared light. These semiconductors enrich the category of plasmonic materials significantly, making it feasible to design plasmonic semiconductor photocatalysts for ammonia synthesis.
![]() | ||
| Fig. 9 (a) Synthesis of the Au/HCNS-NV. (b) NH3 photocatalysis over HCNSx-NV. (c) 1H NMR spectra of the reaction solution upon the photocatalytic N2 fixation reaction for 2 h using Au/HCNS-NV as the catalyst in Ar, 14N2 and 15N2 atmospheres, respectively. (d) Adsorption configuration and charge density difference of N2 adsorbed on the HCNS-NV. Reproduced with permission.73 Copyright 2020, The Royal Society of Chemistry. (e) Nitrogen vacancy generation energies of different samples. (f) EPR spectra of different samples. (g) NH3 yield for different samples with different concentrations of nitrogen vacancies. Reproduced with permission.74 Copyright 2020, Wiley-VCH. | ||
The shell thickness was another critical factor for the performance of photocatalysis as it determined the transport ability of the charge carriers while affecting the interplay of the total number of active sites on the surface and in the mesopores with ease of access. The highest productivity of ammonia was observed on carbon nitride with a shell thickness of 64 nm under visible-light illumination (Fig. 9b). The isotope labelling experiments showed that ammonia originates from nitrogen gas rather than carbon nitride (Fig. 9c). The reaction mechanism was further elucidated by DFT calculations. The nitrogen vacancies at the N site in HCNs-NV captured and held N2 molecules through a di-nuclear end-on coordination mode. The charge density difference indicated that the localized electrons at the nitrogen vacancy sites cause a transfer of electrons to the adsorbed N, weakening the N
N bond (Fig. 9d). Simultaneously, plasmonic Au nanoparticles generate hot electrons under illumination, which are then transferred to the active vacancy sites for N2 activation. The synergistic promotion of Au nanoparticles and nitrogen vacancies in the optimal Au/HCNS-NV composition leads to a high ammonia production rate.
As nitrogen vacancies play a crucial role in the photocatalytic ammonia synthesis, increasing the concentration of nitrogen vacancies is important to further improve the catalytic performance. Recently, Wu et al. conducted a study on Au-loaded g-C3N4 nanosheets (AuCNNVs), which demonstrated a high concentration of nitrogen vacancies. The results showed that Au NPs significantly promoted the generation of nitrogen vacancies, which was in line with the calculated nitrogen vacancy generation energy (Fig. 9e).74 The electron paramagnetic resonance (EPR) spectroscopy results in Fig. 9f indicated that AuCNNVs had the highest concentration of nitrogen vacancies among the prepared catalysts. Moreover, Fig. 9g shows a positive correlation between the concentration of nitrogen vacancies and NH3 yield. The LSPR effect of Au also improved the efficiency of visible light utilization in addition to the contribution of nitrogen vacancies. The integration of nitrogen vacancies and the LSPR effect of Au afforded a high ammonia yield of 93 µmol g−1 without any sacrificial agent under visible-light illumination.
![]() | ||
| Fig. 10 (a) Schematic illustration of optimum AuM surface alloys. Reproduced with permission.75 Copyright 2016, American Chemical Society. (b) The mechanism for plasmonic photocatalytic ammonia synthesis over the CuFe alloy. (c) NH3 production rate over different CuFe alloys. (d) NH3 production rate of Cu96Fe4 for 10 cycles. Reproduced with permission.76 Copyright 2020, American Chemical Society. | ||
Compared to plasmonic noble metals like Au, Ag, or Ru, Cu has garnered significant interest due to its substantial LSPR effect in the visible light region and relative affordability. However, Cu nanoparticles are susceptible to corrosion in air and moisture, leading to challenges in the development of Cu-based catalysts with chemical stability. Recently, Hou et al. reported porous CuFe catalysts with varying Cu/Fe ratios by selectively etching Fe elements from Cu21Fe79.76 In this catalyst, the active Fe species for ammonia synthesis were promoted by plasmonic Cu, while surface Fe atoms acted as active sites for N2 adsorption and activation (Fig. 10b). The optimal catalyst, Cu96Fe4, exhibited an ammonia production rate of 342 µmol g−1 h−1, exceeding other reference catalysts under the same conditions (Fig. 10c). Furthermore, Cu96Fe4 retained ∼100% of its initial activity after 10 cycles, implying excellent stability (Fig. 10d). The reaction's linear dependence on light intensity suggested that the process was driven by plasmonic hot electrons. This study represents a novel approach towards developing plasmonic photocatalysts for ammonia synthesis that are both stable and cost-effective using a Cu-based system.
Recently, Li et al. reported a novel photothermal plasmonic catalysis system that excludes nonthermal plasmonic effects such as the hot carrier-driven process.51 In this system, light-induced thermal gradients primarily contributed to the activity enhancement of ammonia synthesis. The photocatalytic process is illustrated in Fig. 11a. A 3 mm-thick Ru–Cs/MgO catalyst was loaded into the reactor. The LED was positioned above the catalysts while a heating block beneath maintained the desired temperature. Under dark conditions, a positive thermal gradient (∇T = T2 − T1, where T1 is the top-surface temperature and T2 is the bottom-surface temperature) formed, whereas illumination caused the photoheating effect to generate a negative gradient (Fig. 11b). Under combined light and heat, the NH3 production rate for the photo-thermal process (4464 µmol g−1 h−1, at Te = 333 °C and ∇T = −184 °C) was consistently higher than under dark thermal conditions (1530 µmol g−1 h−1 at Te = 333 °C and ∇T = +58 °C) at the same temperature (Fig. 11c). Notably, the NH3 production rate initially decreased slightly but then significantly increased with the shift from positive to negative thermal gradients (Fig. 11d). These findings underscore the critical role of negative thermal gradients in the ammonia synthesis reaction.
![]() | ||
| Fig. 11 (a) Schematic illustration of the reaction chamber for ammonia synthesis. (b) Thermal gradients under dark thermal and heated white light illumination. (c) NH3 production rate under dark and illuminated conditions. (d) NH3 production rate as a function of the thermal gradient. (e) Wavelength-dependent ammonia synthesis rates. (f) NH3 production rate as a function of the thermal gradient for Ru–Cs/MgO and Ti2O3 on Ru–Cs/MgO. Reproduced with permission.51 Copyright 2019, American Chemical Society. | ||
In general, molecules tend to move from the regions of high temperature to the regions of low temperature owing to thermophoretic forces. In the case of ammonia synthesis, a negative temperature gradient facilitates the adsorption and activation of reactants, such as N2 and H2, in a higher temperature region (T1), while the product NH3 moves to a lower temperature region (T2) to prevent reverse decomposition. This negative gradient achieves a balance of high reaction activity and conversion yields. Interestingly, the illumination with different wavelengths has a similar effect on ammonia synthesis, unlike plasmonic photocatalysis driven by hot carriers (Fig. 11e). To confirm that the enhancement in ammonia synthesis is not due to nonthermal effects such as hot carrier generation, a controlled experiment was conducted under indirect illumination using Ti2O3, a black photothermal material, on top of the catalysts to absorb all light and convert it to heat. The photothermal heating produced a negative gradient and provided similar enhancements in ammonia synthesis compared to direct illumination over Ru–Cs/MgO (Fig. 11f). Table 1 summarizes the plasmon-enhanced photocatalytic nitrogen reduction performance. It is obvious that Ru–Cs/MgO shows high activity towards photocatalytic ammonia synthesis, which is comparable to those catalysts employed in the thermal catalytic process. This observation confirms that photothermal heating dominates the ammonia synthesis enhancement rather than hot carrier generation. This study offers a universal and scalable strategy for catalysing various exothermic chemical reactions.
| LSPR effects | Catalysts | T (°C) | Light source & bandwidth | Calibrated light intensity (W cm−2) | Catalyst mass (g) | Reaction phase | Reagents | Flow rate (mL min−1) | 15N2 confirmation | NH3 yield (µmol g−1 h−1) | Method(s) of NH3 quantification | AQE | STA | Ref. |
|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
| Hot carrier injection | Au/TiO2-OV | RT | Xe lamp (visible) | n.a. | 0.1 | Liquid | N2 + H2O | 50 | No | 78.6 | Indophenol-blue | 0.82% (at 550 nm) | n.a. | 53 |
| TiO2/Au/a-TiO2/ | 25 | Xe lamp (420 nm filter) | 0.1 | n.a. | Liquid | N2 + H2O | 20 | No | 0.01 | Indophenol | n.a. | n.a. | 54 | |
| TiO2/Au/K+ | 25 | Xe lamp (UV-vis) | 0.1 | n.a. | Liquid | N2 + H2O | 60 | No | 1020 | Nessler's reagents and indophenol-blue | 0.93% (at 550 nm) | n.a. | 55 | |
| CuFe | 25 | Xe lamp (full spectrum) | 0.25 | 0.01 | Liquid | N2 + H2O | Sealed (3 bar) | No | 342 | Nessler's reagents, indophenol-blue and ion chromatography | 0.13% (at 550 nm) | 0.06% | 76 | |
| Au/end-CeO2 | 25 | Diode laser (808 nm) | 8 | 0.001 | Liquid | N2 + H2O | 5 | No | 114 | Indophenol-blue | n.a. | n.a. | 59 | |
| Ti3C2Tx/TiO2 | 25 | Xe lamp (full spectrum) | 0.25 | 0.01 | Liquid | N2 + H2O | Sealed (1 bar) | No | 422 | Indophenol-blue and ion chromatography | 0.07% (at 740 nm) | n.a. | 68 | |
| MoO3−x | 25 | Xe lamp (full spectrum) | n.a. | 0.05 | Liquid | N2 + H2O | Sealed (1 bar) | No | 328 | Nessler's reagents, and ion chromatography | 0.31% (at 808 nm) | n.a. | 69 | |
| SrMoO4 | 20 | Xe lamp (full spectrum) | n.a. | 0.02 | Liquid | N2 + H2O | 100 | No | 3.9 | Indophenol-blue | 0.022% (at 500 nm) | n.a. | 70 | |
| Au/HCNS-NV | 25 | Xe lamp (full spectrum) | 0.1 | 0.05 | Liquid | N2 + H2O | 50 | No | 783 | Nessler's reagents | 0.64% (at 550 nm) | 0.03% | 73 | |
| AuCNNVs | 25 | Xe lamp (visible) | n.a. | 0.05 | Liquid | N2 + H2O | Sealed (1 bar) | No | 184 | Nessler's reagents | n.a | n.a. | 74 | |
| Photothermal effect | Ru–Cs/MgO | 333 | Blue LED | 4.7 | 0.02 | Gas | N2 + H2 | 75 | No | 4464 | Quadrupole mass spectrometer | n.a | n.a. | 51 |
| α-Fe-110s | 413 | Xe lamp (full spectrum) | 6.12 | 0.013 | Gas | N2 + H2 | 60 | Yes | 1260 | Ion chromatography | n.a | n.a. | 78 | |
| Ru/C | 350 | Xe lamp (full spectrum) | 5 | 0.05 | Gas | N2 + H2 | 30 | No | 1750 | Nessler's reagents, and ion chromatography | n.a | n.a. | 79 | |
| K/Ru/TiO2−xHx | 360 | Xe lamp (full spectrum) | n.a. | 0.1 | Gas | N2 + H2 | 6 | No | 113 | Nessler's reagent | n.a | n.a. | 80 | |
| Pt1-Ptn-TiN | 204 | Blue LED (465) | 0.6 | 0.1 | Gas phase | N2 + H2 | 6 | No | 500 | 1H-NMR | 0.03% (at 465 nm) | n.a. | 81 | |
| Near field enhancement | AuRu0.31 | 25 | Xe lamp (full spectrum) | 0.4 | 0.0002 | Liquid | N2 + H2O | Sealed (2 bar) | Yes | 101 | Ion chromatography | 0.017% (at 550 nm) | n.a. | 82 |
| Ru-Vs-CoS/CN | 25 | Xe lamp (full spectrum) | 0 | 0.025 | Liquid | N2 + H2O | Sealed (1 bar) | No | 438 | Indophenol-blue | 1.28% (at 400 nm) | 0.04% | 83 |
N triple bond (941 kJ mol−1).77 The plasmonic materials enable the cleavage of the N
N bond through indirect and direct hot carrier transfer into the adsorbate.30 The former method involves hot carriers transferring to the lowest unoccupied molecular orbital of the adsorbate on the surface of catalysts, while the latter involves direct hot carrier transferring from plasmonic materials to the hybridized metal–adsorbate interface. Different from the indirect mechanism that takes place after the generation of hot carriers, the direct mechanism occurs at the process of plasmon excitation. However, the direct pathway is uncommon in many reactions because the surface orbital hybridization requires an intense metal/adsorbate interaction.
An intriguing study by Xiong et al. demonstrated photocatalytic nitrogen fixation in pure water under mild conditions, highlighting the important role of direct energy transfer in AuRu core-antenna nanostructures.82 In situ Near-Ambient Pressure X-ray Photoelectron Spectroscopy (NAP-XPS) analysis revealed a N 1s peak attributed to atomic nitrogen (Fig. 12a–d), while no spectroscopic signatures for associative intermediates (e.g., *N2Hx) were detected by Diffuse Reflectance Infrared Fourier Transform Spectroscopy (DRIFTS) (Fig. 12e). Additionally, 1H Nuclear Magnetic Resonance (1H NMR) spectra using 15N2 confirmed that the produced NH3 originated from N2 gas (Fig. 12f). Based on these results, the authors proposed that the reaction proceeds via a dissociative mechanism. First-principles calculations indicated that photo-induced plasmonic near-fields play a crucial role in N2 activation. According to the calculations, hot electron transfer caused only minimal N
N bond elongation, while applying a local electric field resulted in significant bond stretching (Fig. 12g). Combining the effects of hot electrons and local electric fields, a proposed reaction mechanism suggests that the N2 molecule is chemisorbed at Ru sites to form a highly hybridized Ru–N2 complex, facilitated by the local electric field generated from plasmonic Au excitation. This study emphasizes the potential contribution of the plasmonic near-field to N2 activation. However, definitive evidence for a dissociative mechanism in this system requires further substantiation, as the inability to detect associative intermediates could be due to their low surface coverage or transient nature.
![]() | ||
| Fig. 12 (a)–(d) N 1s XPS spectra of AuRu0.31. (e) In situ DRIFTS spectra recorded for N2 + H2O over AuRu0.31 under irradiation conditions. (f) 1H NMR spectra of solution after the N2 fixation reaction in the 15N2, 15N2 + 14N2, or 14N2 atmosphere. (g) Optimized structures of N2 adsorbed on the AuRu cluster. Reproduced with permission.82 Copyright 2019, American Chemical Society. (h) Adsorption energies of N2 on different sites. FDTD electric field distribution. (i) FDTD electric field distribution at Ru/CoSx nanoparticle (800 nm irradiation) observed from γ-axis (vertical to incident light). Reproduced with permission.83 Copyright 2019, Wiley-VCH. | ||
Future investigations utilizing time-resolved spectroscopy and isotope scrambling experiments with 14N2/15N2 mixtures could provide more conclusive mechanistic evidence. Nonetheless, this work underscores the significance of local electric fields and presents a novel platform for designing advanced plasmonic catalysts for ammonia synthesis.
In many cases, N2 adsorption occurs through a terminal end-on mode in ammonia synthesis, in which the N atom solely binds with active metal atoms to accept electrons. However, recent studies have shown that a bridging adsorption mode is more favourable for N2 activation in homogeneous catalysis.83 For example, a heterojunction of Ru/CoSx in g-C3N4 nanosheets was reported to achieve N2 activation through a bridging adsorption mode. In this system, the N atoms were adsorbed to the Ru–Co centre with a lower adsorption energy than end-on adsorption to a Ru or Co site, or a S vacancy (Fig. 12h). Such side-on bridging adsorption contributed to the polarization of N
N bonds. Simultaneously, the Schottky barrier between Ru and CoSx endows the interface with electron transfer from CoSx to Ru, and the S vacancies of CoSx can broaden the light absorption range and serve as additional electron donors for light-excited electrons to promote N2 activation. In addition, the local electric field enhanced by the LSPR derived from plasmonic Ru nanoparticles accelerates charge separation and transfer, leading to more efficient N2 activation (Fig. 12i). This work emphasizes the importance of controlling the N2-coordination environment for the improvement of ammonia synthesis performance.
| Catalytic systems | Plasmonic photocatalysis | Thermal catalysis | Plasma catalysis | Electrochemical NRR |
|---|---|---|---|---|
| Energy efficiency | Very low | Moderate to high | Low | Very low |
| Space-time yield | Low | High | Low to moderate | Low |
| Scalability | Limited | Excellent | Limited | Limited |
| Technology readiness | Low | High | Low | Low |
Morphology is also a key factor in determining the LSPR effects of plasmonic catalysts. Previous studies have shown that nanoparticle shape can significantly influence the region of light absorption. However, current research has mainly concentrated on non-anisotropic structures. Nanoparticles with sharp, flat, or elongated geometries, such as bipyramids or branched nanostructures, are known to generate hot electrons and holes more efficiently. Designing plasmonic structures with specific shapes is therefore an important direction for future research. Moreover, the areas of strongest field enhancement often correspond to crystallographically unstable regions that are susceptible to structural reconstruction or deactivation under reaction conditions. As a result, morphological design must consider not only the spatial distribution and dynamic evolution of electromagnetic hotspots and catalytic active sites but also the structural stability of the nanoparticles.
Given these considerations, future research should focus on creating standardized libraries of plasmonic nanocrystals, including nanorods, bipyramids, core–shell structures, and gap antennas, with precisely controlled geometric parameters such as dimensions, aspect ratios, and tip curvatures. Through systematic optical characterization and catalytic testing of these libraries, we can more accurately determine how specific geometric features influence hot carrier energy distribution, local electric field profiles (both intensity and spatial localization), and photothermal conversion efficiency.
As discussed in Section 3, several recent examples illustrate advanced strategies in hybrid design. For instance, site-selective growth of CeO2 at the ends of Au nanorods creates a spatial configuration that separates the reduction site (CeO2 with oxygen vacancies) from the oxidation site (Au), markedly enhancing charge separation. Similarly, Au nanoparticles encapsulated in hollow carbon nitride within core–shell structures allow the plasmonic core to function as an embedded antenna, while the shell offers active sites and protection. Bimetallic plasmonic cores (e.g., AuRu) with tailored electronic structures further demonstrate how optimizing the plasmonic component itself improves hot carrier generation and transfer. These examples show a shift from simple hybridization to precise nanoscale engineering of geometry and the electronic structure. Looking forward, several key points should be considered for the design of plasmonic materials.
(1) Precision in interface and alloy engineering: Future effort must go beyond bulk heterojunctions. Atomic-level control over the interface structure, achieved through techniques like atomic layer deposition, and the design of alloys or intermetallic compounds will allow for fine-tuning the Schottky barrier height and the dynamics of hot carrier injection.
(2) Expanding the material varieties beyond oxides: While metal-oxide semiconductors have dominated, the exploration of hybrids with non-oxide materials, such as nitrides, phosphides, sulfides, and carbides, presents a vast and untapped opportunity. These materials often possess superior electrical conductivity, narrower bandgaps, and different types of vacancies (e.g., N, S, P, and C vacancies) that could lead to novel adsorption and activation pathways for N2.
(3) Integration with theory and data science: the multi-combination of materials, morphologies, and interfaces is immense. The integration of multi-scale theoretical simulations with machine-learning-guided catalyst design will be indispensable for identifying the most promising hybrid configurations and accelerating the discovery process.
Catalyst deactivation is a complex process affected by photochemical corrosion, thermal sintering or reconstruction, and surface poisoning. Strategies to enhance robustness include:
(1) Encapsulating the plasmonic metal core with an inert yet conductive or mesoporous shell (e.g., silica, titania, carbon, etc.) is a promising strategy. An ideal shell should: (i) physically isolate the core from the harsh reaction environment, preventing corrosion and sintering; (ii) allow free diffusion of reactants and products; and (iii) effectively passivate high-surface-energy facets without severely compromising the charge transfer efficiency.
(2) Constructing alloys or intermetallic compounds between the active plasmonic metal (e.g., Cu) and a more stable metal (e.g., Au, Pd, Pt, etc.) can intrinsically enhance the stability from both electronic and thermodynamic perspectives. The alloying component not only modulates the electronic states to optimize reaction pathways but also raises the activation barrier for corrosion reactions. For example, constructing AuCu alloys can effectively suppress Cu leaching while potentially maintaining or even enhancing catalytic performance through synergistic effects.
(3) Selecting reducible supports (e.g., TiO2−x, MoO3−x, etc.) that can form strong interactions with plasmonic nanoparticles, or those that induce an SMSI effect under reaction conditions, can pin the nanoparticles and prevent their migration and sintering. In some cases, an ultra-thin overlayer formed on the nanoparticle surface via SMSI can provide protection without completely blocking the reaction.
To decipher morphological and structural dynamics, such as light-induced sintering, or facet evolution of plasmonic nanoparticles under reactive environments, techniques like Environmental Transmission Electron Microscopy (ETEM) or Optically coupled TEM (OTEM) are indispensable.84 They provide direct, real-time visual evidence at near-atomic resolution, revealing how the catalyst's physical structure adapts and potentially degrades during operation. For uncovering surface chemistry and reaction intermediates, methods such as Near-Ambient Pressure X-ray Photoelectron Spectroscopy (NAP-XPS) and operando Fourier-Transform Infrared (FTIR) spectroscopy are paramount. NAP-XPS can directly identify the chemical states of catalyst surfaces and adsorbed species, while operando FTIR tracks the evolution of key reaction intermediates, together with clarifying the reaction pathway and active sites.
Critically, understanding the plasmonic enhancement itself requires probing the local electromagnetic field and obtaining molecular fingerprints at the nanoscale. This is the exclusive domain of Surface- and Tip-Enhanced Raman Spectroscopy (SERS/TERS). These techniques provide dramatically enhanced vibrational signals, allowing them to probe molecular adsorption and reaction pathways precisely at the electromagnetic “hot spots” directly correlating enhanced fields with catalytic activity. Finally, hot carriers occur on femtosecond to picosecond timescales, and they cannot be observed by conventional methods. Ultrafast transient absorption spectroscopy is an effective tool capable of directly tracking these processes, quantifying the efficiency of initial charge carrier separation and injection, which is the very foundation of plasmon-mediated photocatalysis.
Theoretical research must keep pace with these experimental advances. Traditional Density Functional Theory (DFT), while invaluable for modeling ground-state chemistry and adsorption energies, is inherently inadequate for describing the non-equilibrium, excited-state processes that define plasmonic catalysis. Its limitations in dealing with excited states and strong electronic correlations can lead to spurious conclusions. Therefore, the development and application of higher-level, multi-reference quantum chemical methods, such as emerging theories such as multiconfigurational n-electron valence second-order perturbation theory (e-NEVPT2) and embedded complete-active-space second-order perturbation theory (emb-CASPT2), are critical.85,86 These methods can more accurately describe the excited potential energy surfaces and charge transfer states involved in plasmon-driven reactions, providing a reliable theoretical framework to interpret complex operando data and guide catalyst design.
Compared to thermal catalysis for ammonia synthesis, plasmonic photocatalytic processes display significantly lower activity levels, making accurate detection of NH3 yield challenging due to various pollution sources. To tackle this issue, we provide a “Decision tree for artifact exclusion” box containing practical steps (Fig. 13).
Before testing, rigorous controlled experiments are essential. Catalytic performance should be evaluated in the dark, without a catalyst, in an Ar atmosphere without N2 and H2 to rule out false positives, including ambient NH3/NOx, nitrogen leaching from supports and sacrificial reagents. When NH3 production exceeds the background level, conducting a 15N isotope experiments is strongly recommended to eliminate pollution sources. Additionally, the current method for detecting the NH3 yield primarily relies on spectrometric techniques, specifically Nessler's reagent. Future efforts should include multi-method approaches for NH3 quantification, facilitating comparisons with reported catalysts. For nitrogen-containing catalysts, performing post-reaction XPS or Electron Energy Loss Spectroscopy (EELS) is crucial to rule out lattice nitrogen consumption. Only after passing these verification steps can the reported activity be deemed credible.
For photocatalytic systems, inconsistencies in reaction parameters, such as illumination source, wavelength range and intensity, solution pH, catalyst loading, and sacrificial oxidants, present significant obstacles to comparative studies. Thus, there is an urgent need to develop standard reactors for photocatalysis. These reactors must be designed to ensure a uniform and well-defined light source, incorporate effective heat-sinking to separate genuine thermal effects from bulk photothermal heating, and enable precise control over mass transport. Simultaneously, a strict set of parameters should be mandated for all publications. Standardizing experimental procedures and performance metrics is essential for enhancing data reliability, thereby paving the way for the development of nitrogen fixation demonstrators, pilot plants, and ultimately solar ammonia refineries.
| This journal is © The Royal Society of Chemistry 2026 |