Next Article in Journal
Exploitation of Lignocellulose Fiber-Based Biotemplates to Improve the Performance of an Immobilized TiO2 Photocatalyst
Next Article in Special Issue
Towards Computer-Aided Graphene Covered TiO2-Cu/(CuxOy) Composite Design for the Purpose of Photoinduced Hydrogen Evolution
Previous Article in Journal
Binder-Free Construction of a Methanol Tolerant Pt/TiO2/Carbon Paper Anode by Atomic Layer Deposition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress in Plasmonic Hybrid Photocatalysis for CO2 Photoreduction and C–C Coupling Reactions

1
KU-KIST Graduate School of Converging Science and Technology, Korea University, 145 Anam-ro, Seongbuk-gu, Seoul 02841, Korea
2
Laboratory for Advanced Molecular Probing, Eco-Friendly New Materials Research Center, Korea Research Institute of Chemical Technology, 141 Gajeong-ro, Yuseong-gu, Daejeon 34114, Korea
3
School of Chemical Engineering, Sungkyunkwan University (SKKU), Suwon 440-746, Korea
*
Authors to whom correspondence should be addressed.
Catalysts 2021, 11(2), 155; https://doi.org/10.3390/catal11020155
Submission received: 18 December 2020 / Revised: 13 January 2021 / Accepted: 15 January 2021 / Published: 22 January 2021
(This article belongs to the Special Issue Novel Photocatalysts for Environmental and Energy Applications)

Abstract

:
Plasmonic hybrid nanostructures have been investigated as attractive heterogeneous photocatalysts that can utilize sunlight to produce valuable chemicals. In particular, the efficient photoconversion of CO2 into a stable hydrocarbon with sunlight can be a promising strategy to achieve a sustainable human life on Earth. The next step for hydrocarbons once obtained from CO2 is the carbon–carbon coupling reactions to produce a valuable chemical for energy storage or fine chemicals. For these purposes, plasmonic nanomaterials have been widely investigated as a visible-light-induced photocatalyst to achieve increased efficiency of photochemical reactions with sunlight. In this review, we discuss recent achievements involving plasmonic hybrid photocatalysts that have been investigated for CO and CO2 photoreductions to form multi-carbon products and for C–C coupling reactions, such as the Suzuki–Miyaura coupling reactions.

1. Introduction

Global population growth and industrial development have been continuously causing the consumption of fossil fuels, resulting in environmental pollution and energy shortages. Among various alternative energy sources, sun light is an eco-friendly, clean, and sustainable energy source that can produce more than tens of thousands of terawatts from the Earth’s surface [1,2]. The sun light can be utilized in various ways in producing electricity, photochemical synthesis in plants, and a giant heat source to maintain biological systems on Earth [3,4,5,6,7]. Therefore, utilizing sun light as an energy source has been an attractive research topic in chemistry and materials sciences. Among many applications with light, mimicking the photosynthetic system with a catalyst and sun light will be greatly attracting and challenging areas.
In this regard, the efficient conversion of CO2 into hydrocarbon with sun light is one plausible way to reduce the amount of CO2 at atmosphere by producing stable chemicals, which can be utilized as an energy source when necessary [8,9]. To achieve this challenging goal, the first step is an efficient conversion of CO2 into hydrocarbon, either C1 or C2. The next step is carbon–carbon coupling to produce multi-carbon products. Although these two types of reactions have different aspects, the reactions have the commonality of not only forming carbon-to-carbon bonds, but also their utility in the production of materials that can be used in other fields.
The first developed method for CO2 conversion is the electrochemical approach, traced back to the 19th century [10]. The electrochemical method has the advantages of flexibility in the design of devices and individual optimization of components, but still has the disadvantage of requiring external energy (electricity). Subsequently, the photocatalytic approach was proposed, which was traced back to the 1970s [11,12,13,14]. Inspired by photosynthesis, solar-driven reduction of CO2 has been considered as one of several possible solutions because it is renewable and eco-friendly energy that can reduce CO2 concentration [12,15,16,17]. The general concept of a CO2 reduction process in the photocatalytic system is shown in Figure 1a.
The conversion of CO2 into C2 hydrocarbon is one type of a C–C coupling reaction [9,18]. However, the reaction requires a multi-step reaction involving multiple electrons. It is more difficult to conduct than the simple CO2 conversion reaction. Therefore, the understanding and development of suitable catalysts, reaction conditions, etc. are essentially required to improve the efficiency of CO2 conversion to multi-carbon products.
Carbon–carbon cross-couplings and related reactions, which are the other types of C–C coupling, present an important research direction in the field of chemistry [19,20,21,22,23]. After the reports of C–C coupling reactions in the 1970s [24,25,26,27], the reactions have been regarded as very powerful methods for forming C–C and C–heteroatom bonds. In recognition of their developments and applications conducted in the 1990s [28,29,30,31], in 2010, R. F. Heck, E.-I. Negishi, and A. Suzuki were awarded the Nobel Prize in Chemistry [32]. Even now, studies including the development of catalysts and reaction conditions are being extensively conducted to increase the efficiency of several named reactions, called Suzuki–Miyaura, Heck, Sonogashira, Stille, and Negishi, which are summarized in Figure 1b. The widely accepted mechanism of C–C cross-coupling reactions, including the Suzuki–Miyaura coupling, consists of three steps: (1) the oxidative addition of a catalyst such as palladium to the halide, which is the rate-determining step in most cases, (2) transmetalation, which is an organometallic reaction where the ligands are transferred from one species to the metal (II) complex, (3) the reductive elimination of corresponding products, and the restoration of the palladium catalyst [33]. Sufficient energy and specific reaction conditions are required to overcome the activation energy barrier, transfer the electrons, and make the reaction proceed [32,34,35]. These problems have led to a demand for sustainable, safe, and environmentally-friendly sources, such as solar energy.
Materials that are responsive to sunlight include plasmonic nanomaterials, semiconductors, and photosensitizers. When light is irradiated on the materials, electrons or energies are excited, causing chemical reactions and the transformation of solar energy into chemical energy. Accordingly, it is important to select the appropriate materials that can improve the catalytic efficiency. In particular, visible/IR-light-responsive materials need to be used because the visible-to-IR light accounts for ca. 95% of the solar light, while the proportion of UV of solar light is only ca. 5% [36,37,38,39,40]. Among possible materials, plasmonic nanoparticles, such as Au, Ag, and Cu, show strong interactions with visible-light and localized electromagnetic field due to their localized surface plasmon resonances (LSPR) [41,42,43]. The collective oscillation of electrons by LSPR induces to yield energetic electrons, called hot electrons, which can help boost the chemical reactions [41,44,45]. The changes in the size, shape, and composition of plasmonic nanomaterials can cause interactions in the near-infrared (NIR) region [46,47,48,49,50]. However, the hybridization of plasmonic nanomaterials with other materials such as semiconductors is necessary because of the extremely short lifetime of the hot electrons (<100 fs) [51,52]. In this review, we focus on the recent development of plasmonic nanomaterial-based photocatalysts for CO2 reduction and C–C coupling.

2. Plasmonic Hybrid Photocatalysts for CO2 Reduction into Hydrocarbon with Multi-Carbon Products

The increased CO2 emission is a global problem, which strongly required us to start the immediate reduction [53,54,55]. Accordingly, CO2 conversion into stable chemicals can be one of the key solutions, which can both reduce the amount of CO2 and produce sustainable energy sources [8,56,57,58]. However, CO2 is one of the most thermodynamically stable molecules due to its strong C=O double bond, which has made it difficult for many researchers to convert CO2 into other fuels [37,57,59,60,61]. Even producing multi-carbon products is significantly more difficult than producing single carbon products because greater energy and a complex multi-step, multi-electron transfer processes are required [8,9,60,62]. Nonetheless, the production of multi-carbon chemicals is more desirable because of their higher energy densities, broader applicability, and use as more convenient storages and transportations [9,13,60,63,64,65].
Typically, semiconductors have been used for CO2 conversion as a photocatalyst, in which the electrons are mainly derived from the excitons of photo-induced semiconductors [13,57,66,67,68]. However, the use of the limited wavelength of light have brought about the introduction of plasmonic metal NPs [9,13,69,70,71]. In the following part, we will discuss the introduction of plasmonic hybrid nanomaterials as a photocatalyst to improve the efficiency of CO2 reduction reactions.

2.1. Gold Nanoparticle (AuNP)-Assisted Plasmonic Photocatalysts for CO2 Reduction to Multi-Carbon Products

The gold nanoparticle (AuNP) is one of the promising visible-light-responsive materials that exhibit strong LSPR phenomenon with visible light [72]. Many advantages of AuNPs, including stability, low toxicity, and optical properties, have led many researchers to use AuNPs as photocatalysts [72,73]. There have been attempts to use AuNPs together with other materials, such as carbon nanoparticles and reduced graphene oxide to convert CO2 into formic acid, but not to produce multi-carbon products [69,74].
The formation of more valuable multi-carbon products requires multiple electron–hole pairs involved in CO2 reduction [75]. To generate abundant charge carriers, Tu et al. introduced AuNPs, showing a strong LSPR effect, into a TiO2 hollow shell, which were called Au@TiO2 yolk–shell hollow spheres [76]. According to this report, bare TiO2 reduced CO2 to produce CH4 (1.33 μmoL g−1 h−1), but not C2H6, while the Au@TiO2 yolk–shell generated both CH4 and C2H6, indicating the significant role of AuNPs in enhancing the photocatalytic yield and the generation of multi-carbon species with rates of 2.52 μmoL g−1 h−1 to produce CH4 and 1.67 μmoL g−1 h−1 to produce C2H6, respectively, accelerating multiple e/h+ reactions [76].
Yu et al. investigated product selectivity tuned by a light excitation attribute, which is a type of plasmonic control with polyvinylpyrrolidone (PVP)-capped AuNPs (11.8 ± 2.3 nm) [77]. They determined that higher photon energies (shorter wavelength) and flux (light intensity) tend to produce C2H6 rather than CH4, showing C2H6 selectivity (Figure 2) [77]. Specifically, CH4 production rates increased proportionally at higher photon energies, that is, shorter excitation wavelengths, but C2H6 was produced only at higher photon energies at a fixed laser intensity of 150 mW·cm−2 (Figure 2b). In addition, CH4 production rates increased almost linearly with increasing photon flux, i.e., light intensity (~0.5 NP−1 h−1 of turnover frequency (TOF) under 532 nm, >0.9 NP−1 h−1 of turnover frequency under 488 nm), irrespective of the wavelength of light, and C2H6 generation was observed only at 488 nm of wavelength and above 300 mW cm−2 (~0.6 NP−1 h−1 of TOF at 750 mW cm−2), indicating the presence of a threshold intensity for C2H6 production (Figure 2c,d).
Zhao et al. also conducted plasmonic control of CO2 conversion using metal/ZnO photocatalysts (Figure 3) [78]. They revealed that the production of higher levels of hydrocarbons, such as C2H6, is closely related to the coupling of the surface plasmon resonance (SPR) field with the intrinsic inner electric field, enabling the separation of electron–hole pairs and the polarization and activation of absorbed substrates. They found that Au interacts more strongly with semiconductors than Ag and Pd, which alters the molecular pathway of CO2 conversion, resulting in a tremendous change in the selectivity of products by density functional theory (DFT) calculations, electron paramagnetic resonance spectroscopy, and Raman spectroscopy. By putting their results all together, they determined that only Au/ZnO can produce C2H6 with a rate of ~25 μmol g−1 h−1 (Figure 3b), while Ag/ZnO and Pd/ZnO produce CH4 and CO without multi-carbon products (Figure 3c,d) [78].
Nguyen et al. used metal–organic frameworks (MOFs) with AuNPs for photocatalytic CO2 conversion to methanol and ethanol [79]. They observed the effect of Au loading on the Aux@zeolitic imidazolate framework (ZIF-67) for the reaction in which Au10@ZIF-67 showed the highest performance of CH3OH production at a rate of 1623 μmol g−1 h−1, while Au20@ZIF-67 showed that C2H5OH production occurred at a rate of 495 μmol g−1 h−1 (Figure 4a,b), and both CH3OH and C2H5OH products decreased with Au30@ZIF-67, possibly due to the agglomeration of AuNPs. Because the reaction to form C2H5OH requires more electrons than CH3OH, these phenomena are thought to be the result of higher hot electrons that enable the photoreduction and support C2H5OH generation due to higher Au concentrations [79].

2.2. Silver Nanoparticle (AgNP)-Assisted Plasmonic Photocatalysts for CO2 Reduction to Multi-Carbon Products

The silver nanoparticle (AgNP) is also one of the plasmonic materials that can exhibit unique optical properties for photocatalysts under visible light with LSPR, like AuNP [80]. Cai et al. prepared AgClxBr1-x alloy nanocrystals and found that the conduction band level could be affected by varying compositions [81]. The substitution of Cl with Br leads to a narrower band gap due to a negative shift in the conduction band minimum. Accordingly, AgCl0.75Br0.25 exhibited higher photocatalytic efficiency for CO2 reduction into both CH3OH (181 μmol g−1) and C2H5OH (362 μmol g−1) than any other AgClxBr1-x with different compositions. Furthermore, the amplified electric field due to the LSPR of Ag0 species also contributes to the light enhancement by encouraging the photocatalytic reaction. Cai et al. further developed an Ag/AgCl photocatalyst system with a coaxial tri-cubic morphology, called red Ag/AgCl [82]. The enhancement of light harvesting property with red Ag/AgCl was observed, compared to normal AgCl, due to the synergistic effect between metallic AgNPs and the n-type AgCl semiconductor, which is the featured LSPR, Schottky junction, and polarization effect induced by surface plasmons. As a result, the CH3OH and C2H5OH yields and apparent quantum efficiency for the red Ag/AgCl catalysts were 146 and 223 μmol g−1 for 5 h, respectively, in which both are higher than those for normal AgCl catalysts (Figure 5) [82].
Li et al. suggested a system with a plasmonic photocatalyst other than the C–C coupling of CO2 alone by adopting the oxidative coupling of methane using CO2 as the oxidant [83]. In the Ag/TiO2 system, AgNPs can absorb visible light and generate hot electrons and holes, and the hot electrons are injected into TiO2 while hot holes should be captured by CH4. Otherwise, it will lead to the accumulation of Ag(I), which can be reduced back to Ag(0) under UV light. At the same time, TiO2 can generate photoexcited electron–hole pairs, and photoexcited holes can combine with hot electrons from Ag, while photoexcited electrons can reduce CO2 adsorbed on TiO2. This synergistic effect enhances the photocatalytic activity of the reaction (1149 μmoL g−1 h−1 for CO and 686 μmol g−1 h−1 for C2H4) and contributes to the high stability of the catalyst (Figure 6b) [83]. They also demonstrated that other types of support materials with Ag did not present any synergistic effect due to the formation of an unsuitable Schottky barrier and CO2 adsorption sites (Figure 6c), and other precious metals with TiO2 showed lower activity for the photocatalytic reaction due to the poor SPR effect of the metals (Figure 6d) [83].

2.3. Copper Nanoparticle (CuNP)-Assisted Photocatalytic CO2 Reduction to Multi-Carbon Products

Metallic Cu species, including CuNPs, have been used as cocatalysts due to their known effectiveness in generating not only C1 products, but also multi-carbon organic compounds, not as light absorbers with the LSPR effect [84,85,86,87]. Therefore, Cu has been used as an electron acceptor and suppressor of recombination of photoexcited electron–hole pairs generated from semiconductors [88,89,90]. Shown et al. prepared CuNP-decorated graphene oxide because of the large work function of Cu compared to that of GO, producing methanol and acetaldehyde [88]. They controlled the production rates and the ratio of and between both products by adjusting the work function of Cu/GO hybrids [88]. Park et al. used trititanate nanotubes (TNTs) decorated with Cu and CdS quantum dots (CdS/Cu-TNTs) for the production of C1–C3 hydrocarbons [89]. When irradiated with light, an efficient reduction of CO2 to C1–C3 hydrocarbons was observed by the transport of photogenerated electrons to the Cu part through the TNTs while photogenerated holes oxidize water, which is similar to artificial photosynthesis [89]. Chen et al. carried out photocatalytic CO2 reduction with benzyl alcohol oxidation to benzyl acetate using Cu2O/Cu nanocomposites due to the narrow direct band gap and the position of the conduction band of Cu2O [90]. Electrons and holes are generated over Cu2O with visible light irradiation, and the electrons transfer to the surface of Cu due to their lower work function, while the holes can react with benzyl alcohol to form benzaldehyde, followed by the subsequent coupling reaction to form benzyl acetate [90]. CuNPs have great potential for photocatalytic C–C coupling, but more research is needed to utilize CuNPs for light harvesting materials.
There have been several attempts to produce value-added fuels through CO2 conversion using plasmonic hybrid photocatalysts, even though only a few studies have been conducted to generate single carbon products. The hybrid structures of the photocatalysts, reaction conditions, and yields are summarized in Table 1.

3. Plasmonic Hybrid Photocatalysts for C–C Cross-Coupling

The Suzuki–Miyaura coupling, which is one of the most powerful methods for carbon–carbon cross-coupling, is the reaction between an organoboron species and aryl halide in the presence of a palladium (Pd) catalyst and base [21,91]. Although most of this reaction involves a Pd catalyst, which has been regarded as a catalytically active site for the reaction, Pd is difficult to use as a photocatalyst because of its low absorption of visible light [92]. In recent years, to remedy this shortcoming, plasmonic nanoparticles and semiconductors that can absorb visible light have been used. In particular, plasmonic NPs (Au, Ag, Cu) are useful materials that can help the reaction of the Pd catalyst because they interact with visible light to show a strong LSPR phenomenon [93,94,95]. In addition, semiconductors are introduced into plasmonic materials, so-called plasmonic hybrid structures, to help plasmonic NPs extend the lifetime of hot electrons excited by light, or to form electron–hole pairs [96,97]. The hybrid structures classified by each type of plasmonic NPs for photocatalytic C–C cross-coupling reactions are described below.

3.1. AuNP-Assisted Plasmonic Photocatalysts for C–C Cross-Coupling

For C–C cross-coupling reactions, Au–Pd nanocomposites can be utilized, where the Au part absorbs visible light and transfers hot electrons into Pd, and the Pd part acts as electron acceptors and active sites. Some studies using Au–Pd alloys without non-plasmonic materials, such as AuPd nano-wheels [98] and AuPd nanotriangles [99] for photocatalytic C–C cross-coupling, have been reported (Figure 7). Huang et al. prepared AuPd nano-wheels, in which Pd encircles an Au core, with a controllable edge length and tunable SPR using a facile wet-chemical reduction method. In this work, the photocatalytic efficiency of the nano-wheel-catalysts for benzyl alcohol conversion and Suzuki coupling was confirmed, and the yield of products was 65.8% at 50 °C for 1 h under visible light (Figure 7a) [98]. Gangishetty et al. synthesized AuPd bimetallic nanotriangles consisting of an Au nanotriangle core with an unevenly distributed Pd shell, which is similar in morphology with AuPd nano-wheels [99]. The nano-catalysts showed >80% yield of Suzuki coupling between p-iodobenzoic acid and phenylboronic acid for 5 h under a green LED, accompanied by an increase in the temperature from 25 °C to 37 °C while the dark reaction showed only ca. 35% lower conversion compared to the light reaction (Figure 7b). Compared to the yield under dark reaction at 37 °C (ca. 75%), the pure photocatalytic effect is not significant, while the photothermal effect generated from non-radiative plasmon decay is the primary factor (Figure 7b) [99].
Sarina, Xiao et al. prepared Au–Pd alloy NPs embedded on ZrO2, which has a band gap of approximately 5 eV, exhibiting negligible visible light absorption above 400 nm, so that the ZrO2 support does not contribute to the photocatalytic activity of C–C cross-coupling (Figure 8) [100,101,102,103]. Several C–C coupling reactions such as Suzuki–Miyaura, Sonogashira, Stille, Hiyama, Buchwald–Hartwig, and Ullmann coupling, were conducted to study the effects of the wavelength and intensity of the light, the Au/Pd molar ratio of the alloy NPs on the photocatalytic activity, and the photocatalytic mechanism for the C–C coupling reactions. The researchers described the photocatalytic process in view of reaction kinetics. The reduced activation energy for the C–C coupling reaction is possible by visible-light absorption of photocatalyst, indicating the low activation energy in the photocatalytic process compared with that of a thermal reaction process (Figure 8b,c) [101,102]. In terms of energy levels of the molecular orbital, there are two light absorption mechanisms such as inter-band excitation and LSPR absorption at 530 nm for AuNPs. Inter-band excitation is possible with short-wavelength (e.g., 400 nm) absorption via a single-electron excitation. When irradiated with short-wavelength, single-photon excitation generates hot electrons to be injected into the lowest unoccupied molecular orbital (LUMO). In the case irradiated with longer wavelengths, the hot electrons generated by the collective excitation of LSPR can only induce reactions with lower energy thresholds (Figure 8d) [103]. In spite of the low TOF and the number of conversion reactions, they provided a comprehensive insight for the photocatalytic reactions for various C–C cross-coupling reactions and explained the kinetics and mechanisms of the reactions.
Other researchers have used a variety of support materials such as semiconductors (e.g., CeO2, graphitic carbon nitride (g-C3N4)) [104,105], metal–organic frameworks (MOF, UiO-66-NH2) [106], polymers (e.g., perylene bisimide (PBI), polystyrene) [107,108], wide band gap semiconductors (TiO2) [109,110,111,112], and silica [113]. Semiconductors that have narrower band gaps than 3.1 eV (i.e., 2.7 eV for g-C3N4) can absorb visible light to generate electron–hole pairs, which are transferred to metal NPs. By itself, a semiconductor can help the Pd catalyst in photocatalytic C–C cross-coupling reactions [114,115,116], but even the synergy of plasmonic hybrid with semiconductors can be expected to have greater photocatalytic efficiency for a C–C cross-coupling reaction. In these systems, the hot electrons generated from AuNPs (due to the strong LSPR effect of AuNPs) using irradiation can be injected into the attached Pd, in which the electrons transfer to aryl halide molecules, while the photogenerated electron–hole pairs of semiconductors by irradiation can be separated, causing electrons to recover into the Au0 state and holes to transfer into solvent or phenylboronic acid to be activated (Figure 9) [104,105]. Moreover, semiconductors such as g-C3N4 are two-dimensional materials with large surface areas and unique electronic/optical properties, and Schottky junctions form at the interface of the metal and semiconductor, relying on the band alignment and work function [114,115,116], leading to a positive effect on catalytic efficiency (Figure 9b) [105]. Therefore, the AuPd/g-C3N4 nanohybrid showed a very high TOF of 7920 h−1 in the C-C cross-coupling reaction.
MOFs such as UiO-66-NH2 are similar to semiconductors in that they have a band gap (approximately 2.67 eV) capable of absorbing visible light. The key differences are that the energy transfer occurs from ligands to a metal. The pore volume and surface area can be controlled by using the functional group. Noble metals can be introduced into MOFs, and they exhibit high dispersion stability owing to the ultra-high surface area [106].
In general, wide band gap semiconductors such as TiO2 display great light absorption at wavelengths below 400 nm, making it difficult to expect visible light absorption for catalytic reactions [37]. Furthermore, the high energy of UV light might oxidize and reduce organic reactants, which can result in low yield and selectivity. Rohani et al. changed the morphology of TiO2 into urchin-like yolk–shell structures with unique properties, such as a high surface area and visible-light harvesting (Figure 10) [112]. The hydrogenated urchin-like yolk@shell TiO2 structure (HUY@S-TOH) was decorated with plasmonic Au/Pd NPs for photocatalytic Suzuki coupling. The structure showed absorption property for visible light because of the Ti3+ species on the surface of the structure and AuNPs, enhancing the light harvesting efficiency and the inhibition of the recombination of photogenerated electron–hole pairs by decreasing the band gap of TiO2 to the visible region. In addition to the strong interaction between noble metals and TiO2-x, HUY@S-TOH/AuPd showed a high catalytic efficiency with a TOF value of 7095 h−1.
Graphene and its derivatives are 2D materials with an exceptional electron mobility of 2 × 105 cm2·v−1s−1 and variable band gap that depends on the oxidation state of graphene [69,117,118]. Hybrid structures containing graphene or its oxide have been reported for photocatalytic C–C cross-coupling to extend the lifetime of hot electrons of plasmonic NPs generated by light absorption or to prevent recombination of electron–hole pairs [119,120]. Moreover, graphene or slightly oxidized graphene on its own can also be a good support material not only for combination with Pd2+ or Pd0 NPs owing to its functional group, but also for transferring electrons into Pd with facilities to enhance catalytic activity [121]. Kang et al. studied the effect of interfaces on Pd-nanodot-decorated AuNPs with a graphene layer with different oxidation states for photocatalytic Suzuki coupling (Figure 11) [119]. They prepared Pd-cys-AuNPs, Pd-GO-AuNPs, and Pd-rGO-AuNPs with different oxidation states, and found that Pd-rGO-AuNPs exhibited the fastest reaction progress (66.4% with a thermal effect and 54.4% without a thermal effect), while the Pd-cys-AuNPs exhibited the slowest reaction progress (30% with a thermal effect and 6.7% without a thermal effect) for 2 h (Figure 11b,c). The contribution of the electron transfer mechanism from plasmonic NPs (AuNPs) to Pd nanodots is significant through the graphene interface, preventing the relaxation of hot electrons of plasmonic NPs induced by light [119].
Plasmonic properties can be tuned by size and shape by utilizing varying wavelengths of light [46,122]. Gold nanorods (AuNRs), like spherical AuNPs and other noble metal NPs, have the ability to interact with light of varying wavelengths through LSPR [46,123,124,125]. They display two SPR bands of transverse and longitudinal bands due to their anisotropic shape. The transverse mode is located near 500 nm, while the longitudinal mode varies widely depending on the aspect ratio and the overall size of AuNRs, generally located in the NIR region. Therefore, AuNRs have been used as NIR-responsive photocatalysts [122,126].
AuNR-based photocatalysts have been applied to C–C coupling reactions [127,128,129,130,131]. Wang et al. synthesized Au–Pd nanostructures where Pd NPs were located at the tip of the AuNRs to a large extent as well as on the whole surface of the AuNRs (Figure 12a). They conducted photocatalytic Suzuki coupling under an 809-nm laser in the NIR region, yielding biphenyl products, where they demonstrated the two origins of the catalytic activity of Au–Pd nanostructures, which is the plasmonic heating process and plasmon-excitation-induced hot electrons [127]. Guo et al. developed Au–Pd nanostructures, called Au@Pd superstructures formed from different directing agents, investigating the effect of the shapes of Pd, which are superstructures, nano-dendrites, and shell structures (Figure 12b). Different from AuNRs, the Au@Pd core@shell exhibited weak electric field enhancement due to the plasmon shielding effect of Pd shells. In the case of Au@Pd nano-dendrites with discrete Pd surfaces, the |E|/|E0| around their surface was largely enhanced by 16 of the maximum value of |Emax|/|E0|, which allows the inside of the AuNR core to be excited partially. The Au@Pd superstructures displayed further improvement of |Emax|/|E0| up to 23 because of the ordered open structure of the Pd nano-arrays and their strong plasmonic antenna effect. Accordingly, the plasmon-enhanced photocatalytic activity of Au@Pd superstructures was >4 times higher than that of the Au@Pd core@shell (TOF ≈ 2880 h−1), and approximately two times higher than that of the Au@Pd nanodendrites [129]. Yoshii et al. designed Pd-graphene-AuNR nanocomposite catalysts similar to Pd-rGO-Au by Kang et al., facilitating the transfer of SPR-induced hot electrons by AuNR to the catalytic active metals (Pd) through the graphene layer (Figure 12c). Almost all aspects of Pd-graphene-AuNR nanocomposites are similar to Pd-rGO-Au, and the only difference is that AuNR was used to absorb NIR light [131].

3.2. AgNP-Assisted Plasmonic Photocatalysts for C–C Cross-Coupling

AgNP-based hybrid photocatalysts with various support materials such as graphene oxide with AgBr [132], silica [113,133], supramolecular ensemble with Cu2O [134,135], and TiO2 [136] for C–C cross-coupling have been reported. Verma et al. used mesoporous silica as a support material (SBA-15, Aldrich), of which the channel can contain Pd/Ag NPs with stability for photocatalytic Suzuki coupling [113,133]. They observed the effect of the shape of Ag or type of plasmonic NPs (Au or Ag), affecting plasmonic optical properties and catalytic activity (Figure 13). They found that longer aspect ratios of the Ag nanostructure result in higher photocatalytic activity efficiency due to the light absorption property (<30% for Pd/Ag/SBA-15(Y), ~40% for Pd/Ag/SBA-15(R), 53% for Pd/AgSBA-15(B)) (Figure 13b) [133]. Although the photocatalytic activity for Suzuki coupling with Pd/Au/SBA-15 was better than that for Suzuki coupling with Pd/Ag/SBA-15 (~70% for Pd/Au/SBA-15, and ~40% for Pd/Ag/SBA-15) (Figure 13d), the activity with Pd/Ag/SBA-15 for the dehydrogenation reaction was better than that with Pd/Au/SBA-15, indicating that it is difficult to simply compare the effects of those reactions [113]. Putting it all together, it is meaningful that Ag and Au can work complementarily with each other in terms of light absorption.
Bhalla’s group used supramolecular ensembles as both reactors and stabilizers of NPs to a higher extent through electron-rich assemblies and introduced Cu2O as a shell around the AgNP for its affordable price, stability, and property. Cu2O is a p-type semiconductor and has been used as an efficient catalyst for C–C, C–N, and C–O cross-coupling reactions (Figure 14) [134,135]. These reports are not studies for the development of conventional Suzuki coupling, but they are significant in that they were performed without palladium even though they showed relatively low photocatalytic efficiency.

3.3. CuNP-Assisted Plasmonic Photocatalysts for C–C Cross-Coupling

Copper nanoparticles (CuNPs) are plasmonic materials like AuNPs and AgNPs, which possess unique optical properties [137,138,139]. However, their instability and tendency to undergo surface oxidation make it difficult for many researchers to utilize CuNPs [140]. Nevertheless, there have been some attempts to overcome the instability in order to use CuNPs for photocatalysts because of their low cost [141,142,143].
For C–C cross-coupling with CuNPs, CuPd bimetallic alloy NPs with other light-absorbing materials such as silicon carbide (SiC) as a semiconductor and NH2-UiO-66(Zr) as MOF were used under visible light (Figure 15) [144,145]. Wang et al. prepared PdCu/SiC using a sol-gel and carbon thermal reduction process for photocatalytic Sonogashira reaction under visible light. They proposed a mechanism in which photogenerated electrons transfer to the Pd part, facilitating the cleavage of aryl halides, while photogenerated holes in the CuNPs react with phenylacetylene to form a phenylethynylcopper(I) compound (Figure 15a) [144]. Sun et al. prepared CuPd@NH2-UiO-66(Zr), with encapsulated bimetallic CuPd nanoclusters inside the cavities of NH2-UiO-66(Zr) via double-solvent impregnation followed by chemical reduction for a Suzuki coupling reaction. The transfer of the photogenerated electrons from the photoexcited NH2-UiO-66(Zr) to the Pd part to form electron-rich Pd was facilitated by metallic Cu acts as an electron mediator, which results in the superior activity of photocatalytic Suzuki coupling, since Cu has a higher Fermi energy level as compared to Pd and lower Fermi energy level as compared to NH2-UiO-66(Zr) (Figure 15b) [145]. These reports do not state the plasmonic light-harvesting property of CuNP itself as a photocatalyst, but attempts to introduce copper into Pd.
In order to stabilize CuNPs, Cui et al. introduced graphene to Cu because of the possible change of the electronic structure of Cu by the carbon vacancies or dangling bond in graphene [146,147]. Due to the LSPR effect of CuNPs, the electron density in Cu is polarized, causing charge heterogeneity at the surface of CuNPs with both relatively electron-rich sites and positively charged sites. The electron-rich sites can easily adsorb imidazole molecules and inject into the molecules, facilitating the cleavage of N–H bonds, while positively charged sites can assist to cleave C–B bonds in phenylboronic acid molecules, and, as a result, C–N bonds are formed (Figure 16a). Furthermore, graphene can also absorb light, generating a strong photocurrent, and the work function of graphene, which is lower than that of Cu, causes hot electrons to transfer to Cu from graphene easily, which can result in the collection of energetic electrons at the Cu sites to accelerate the reaction [147]. Bhalla et al. used supramolecular ensembles as reactors and stabilizers of CuNPs, which is similar to supramolecular ensemble-based Ag@Cu2O core@shell NPs [148]. They confirmed the existence of the LSPR band of CuNPs for plasmonic photocatalysts, and the hybrid systems showed efficient photocatalytic efficiency for photocatalytic C(sp2)–H alkynylation (another type of C–C coupling reaction) (Figure 16b). These reports are not about the named C–C cross-coupling reactions, but they are meaningful in that they make good use of the plasmonic properties of CuNPs as a photocatalyst.
Many attempts have been made to introduce Pd NPs into solid supports to stabilize or assist Pd for catalytic efficiency of C–C cross-coupling, in addition to the examples mentioned above. The hybrid structures of the plasmonic photocatalysts used, reaction conditions, and yields with TOF values for C–C cross-coupling are classified and summarized in Table 2.

4. Summary and Outlook

Herein, we reviewed the research and development of plasmonic hybrid nano-catalysts for two types of photocatalytic C–C coupling reactions: C–C coupling in CO2 reduction to hydrocarbon fuels and C–C cross-coupling in organic chemistry. The C–C coupling in CO2 reduction into multi-carbon fuels can be promoted with the aid of plasmonic NPs as both light absorbers and affinity sites for reactants. Likewise, Suzuki coupling can also be represented in C–C cross-coupling reactions, including Heck, Sonogashira, Stille, Negishi, and other reactions, typically using Pd as an active catalyst and plasmonic NPs with semiconductors as light-responsive materials to accelerate photocatalytic reactions. As mentioned above, the performances of the photocatalytic CO2 reduction and C–C cross-coupling reactions are summarized in Table 1 and Table 2.
Although plasmonic hybrid nano-catalysts for photocatalytic C–C cross-coupling in CO2 reduction to hydrocarbon and C–C coupling in organic chemistry have been intensively investigated for their ability to contribute to the fine-chemical industry, energy sector, and environmental fields, and for solving challenges that still exist in synthesizing suitable photocatalysts by reaching high quantum yields for commercialization, controlling photothermal effects may affect reactions, and help understand the mechanism. As for C–C coupling in CO2 reduction to value-added products, multi-carbon production is still much more difficult than C1 production because of the requirement of multiple electrons, steps, and low selectivity. In addition, Cu has not been properly utilized as a plasmonic material that is cheaper than gold and silver. Thus, developing new hybrid materials with high efficiency could be one of the possible solutions. As for C–C cross-coupling in organic chemistry, the photocatalytic reaction, even Suzuki coupling, has no unified reaction system, where the efficiency varies widely depending on the reaction conditions, reactants, etc., and integrated mechanism. In addition, only a few studies on other photocatalytic C–C coupling reactions including Heck, Sonogashira, Stille, and Negishi coupling, except Suzuki coupling, have been conducted. Moreover, from an economic point of view, C–C cross-coupling reactions typically require expensive Pd active catalysts, which makes commercialization more difficult, so the Pd-free reaction needs to be actively studied.

Author Contributions

Conceptualization, D.-K.L. Investigation, H.H.S. Writing—Original Draft Preparation, H.H.S. Writing—Review and Editing, Y.D.S. and D.-K.L. Supervision, D.-K.L. Project Administration, Y.D.S. and D.-K.L. Funding Acquisition, Y.D.S. and D.-K.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research is supported by the KRICT (KK2061-23), Bio Industrial Strategic Technology Development Program (No. 10077582) funded by the Ministry of Trade, Industry & Energy (MOTIE, Korea), and the Global Research Laboratory (GRL) Program of the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (No. 2016911815). This research is also supported by the National Research Foundation of Korea (2016R1A2B3013825, 2017M3D1A1039421, 2016M3A9B5940991) and the KU-KIST Research Fund.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Aman, M.M.; Solangi, K.H.; Hossain, M.S.; Badarudin, A.; Jasmon, G.B.; Mokhlis, H.; Bakar, A.H.A.; Kazi, S.N. A review of safety, health and environmental (SHE) issues of solar energy system. Renew. Sustain. Energy Rev. 2015, 41, 1190–1204. [Google Scholar] [CrossRef]
  2. Hosenuzzaman, M.; Rahim, N.A.; Selvaraj, J.; Hasanuzzaman, M.; Malek, A.B.M.A.; Nahar, A. Global prospects, progress, policies, and environmental impact of solar photovoltaic power generation. Renew. Sustain. Energy Rev. 2015, 41, 284–297. [Google Scholar] [CrossRef]
  3. Spasiano, D.; Marotta, R.; Malato, S.; Fernandez-Ibañez, P.; Di Somma, I. Solar photocatalysis: Materials, reactors, some commercial, and pre-industrialized applications. a comprehensive approach. Appl. Catal. B 2015, 170, 90–123. [Google Scholar] [CrossRef]
  4. Long, R.; Li, Y.; Song, L.; Xiong, Y. Coupling solar energy into reactions: Materials design for surface plasmon-mediated catalysis. Small 2015, 11, 3873–3889. [Google Scholar] [CrossRef] [PubMed]
  5. Smith, J.G.; Faucheaux, J.A.; Jain, P.K. Plasmon resonances for solar energy harvesting: A mechanistic outlook. Nano Today 2015, 10, 67–80. [Google Scholar] [CrossRef]
  6. Goodenough, J.B. Energy storage materials: A perspective. Energy Storage Mater. 2015, 1, 158–161. [Google Scholar] [CrossRef]
  7. Halabi, M.A.; Al-Qattan, A.; Al-Otaibi, A. Application of solar energy in the oil industry—Current status and future prospects. Renew. Sustain. Energy Rev. 2015, 43, 296–314. [Google Scholar] [CrossRef]
  8. Alper, E.; Orhan, O.Y. CO2 utilization: Developments in conversion processes. Petroleum 2017, 3, 109–126. [Google Scholar] [CrossRef]
  9. Du, C.; Wang, X.; Chen, W.; Feng, S.; Wen, J.; Wu, Y.A. CO2 transformation to multicarbon products by photocatalysis and electrocatalysis. Mater. Today Adv. 2020, 6, 100071. [Google Scholar] [CrossRef]
  10. Royer, M.E. Réduction de l’acide carbonique en acide formique. Comptes Rendus 1870, 1870, 731–732. [Google Scholar]
  11. Halmann, M. Photoelectrochemical reduction of aqueous carbon dioxide on p-type gallium phosphide in liquid junction solar cells. Nature 1978, 275, 115–116. [Google Scholar] [CrossRef]
  12. Inoue, T.; Fujishima, A.; Konishi, S.; Honda, K. Photoelectrocatalytic reduction of carbon dioxide in aqueous suspensions of semiconductor powders. Nature 1979, 277, 637–638. [Google Scholar] [CrossRef]
  13. Li, K.; Peng, B.; Peng, T. Recent advances in heterogeneous photocatalytic CO2 conversion to solar fuels. ACS Catal. 2016, 6, 7485–7527. [Google Scholar] [CrossRef]
  14. Wang, C.; Sun, Z.; Zheng, Y.; Hu, Y.H. Recent progress in visible light photocatalytic conversion of carbon dioxide. J. Mater. Chem. A 2019, 7, 865–887. [Google Scholar] [CrossRef]
  15. Kim, D.; Sakimoto, K.K.; Hong, D.; Yang, P. Artificial photosynthesis for sustainable fuel and chemical production. Angew. Chem. Int. Ed. 2015, 54, 3259–3266. [Google Scholar] [CrossRef]
  16. Marshall, J. Solar energy: Springtime for the artificial leaf. Nature 2014, 510, 22–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Creel, E.B.; McCloskey, B.D. Scalable CO2-to-oxygenate production. Nat. Catal. 2018, 1, 6–7. [Google Scholar] [CrossRef]
  18. Bushuyev, O.S.; De Luna, P.; Dinh, C.T.; Tao, L.; Saur, G.; van de Lagemaat, J.; Kelley, S.O.; Sargent, E.H. What should we make with CO2 and how can we make it? Joule 2018, 2, 825–832. [Google Scholar] [CrossRef] [Green Version]
  19. Li, H.; Johansson Seechurn, C.C.C.; Colacot, T.J. Development of preformed Pd catalysts for cross-coupling reactions, beyond the 2010 Nobel prize. ACS Catal. 2012, 2, 1147–1164. [Google Scholar] [CrossRef]
  20. Shi, W.; Liu, C.; Lei, A. Transition-metal catalyzed oxidative cross-coupling reactions to form C-C bonds involving organometallic reagents as nucleophiles. Chem. Soc. Rev. 2011, 40, 2761–2776. [Google Scholar] [CrossRef]
  21. Miyaura, N.; Suzuki, A. Palladium-catalyzed cross-coupling reactions of organoboron compounds. Chem. Rev. 1995, 95, 2457–2483. [Google Scholar] [CrossRef] [Green Version]
  22. Suzuki, A. Cross-coupling reactions of organoboranes: An easy way to construct C–C bonds (Nobel lecture). Angew. Chem. Int. Ed. 2011, 50, 6722–6737. [Google Scholar] [CrossRef] [PubMed]
  23. Bariwal, J.; Van der Eycken, E. C–N bond forming cross-coupling reactions: An overview. Chem. Soc. Rev. 2013, 42, 9283–9303. [Google Scholar] [CrossRef]
  24. Tamao, K.; Sumitani, K.; Kumada, M. Selective carbon–carbon bond formation by cross-coupling of grignard reagents with organic halides. catalysis by nickel-phosphine complexes. J. Am. Chem. Soc. 1972, 94, 4374–4376. [Google Scholar] [CrossRef]
  25. Miyaura, N.; Yamada, K.; Suzuki, A. A new stereospecific cross-coupling by the palladium-catalyzed reaction of 1-alkenylboranes with 1-alkenyl or 1-alkynyl halides. Tetrahedron Lett. 1979, 20, 3437–3440. [Google Scholar] [CrossRef] [Green Version]
  26. Heck, R.F.; Nolley, J.P., Jr. Palladium-catalyzed vinylic hydrogen substitution reactions with aryl, benzyl, and styryl halides. J. Org. Chem. 1972, 37, 2320–2322. [Google Scholar] [CrossRef]
  27. Negishi, E.-I.; Baba, S. Novel stereoselective alkenyl–aryl coupling via nickel-catalysed reaction of alkenylanes with aryl halides. J. Chem. Soc. Chem. Commun. 1976, 596b–597b. [Google Scholar] [CrossRef]
  28. Ishiyama, T.; Matsuda, N.; Miyaura, N.; Suzuki, A. Platinum(0)-catalyzed diboration of alkynes. J. Am. Chem. Soc. 1993, 115, 11018–11019. [Google Scholar] [CrossRef]
  29. Ishiyama, T.; Murata, M.; Miyaura, N. Palladium(0)-catalyzed cross-coupling reaction of alkoxydiboron with haloarenes: A direct procedure for arylboronic esters. J. Org. Chem. 1995, 60, 7508–7510. [Google Scholar] [CrossRef]
  30. Guram, A.S.; Rennels, R.A.; Buchwald, S.L. A simple catalytic method for the conversion of aryl bromides to arylamines. Angew. Chem. Int. Ed. Engl. 1995, 34, 1348–1350. [Google Scholar] [CrossRef]
  31. Louie, J.; Hartwig, J.F. Palladium-catalyzed synthesis of arylamines from aryl halides. Mechanistic studies lead to coupling in the absence of tin reagents. Tetrahedron Lett. 1995, 36, 3609–3612. [Google Scholar] [CrossRef]
  32. Johansson Seechurn, C.C.C.; Kitching, M.O.; Colacot, T.J.; Snieckus, V. Palladium-catalyzed cross-coupling: A historical contextual perspective to the 2010 Nobel prize. Angew. Chem. Int. Ed. 2012, 51, 5062–5085. [Google Scholar] [CrossRef] [PubMed]
  33. Diederich, F.; Stang, P.J. Metal-Catalyzed Cross-Coupling Reactions; Wiley-VCH: Weinheim, Germany, 2004. [Google Scholar]
  34. Fihri, A.; Bouhrara, M.; Nekoueishahraki, B.; Basset, J.-M.; Polshettiwar, V. Nanocatalysts for Suzuki cross-coupling reactions. Chem. Soc. Rev. 2011, 40, 5181–5203. [Google Scholar] [CrossRef] [PubMed]
  35. Biffis, A.; Centomo, P.; Del Zotto, A.; Zecca, M. Pd metal catalysts for cross-couplings and related reactions in the 21st century: A critical review. Chem. Rev. 2018, 118, 2249–2295. [Google Scholar] [CrossRef]
  36. Linsebigler, A.L.; Lu, G.; Yates, J.T., Jr. Photocatalysis on TiO2 surfaces: Principles, mechanisms, and selected results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  37. Chen, X.; Mao, S.S. Titanium dioxide nanomaterials: Synthesis, properties, modifications, and applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef]
  38. Zayat, M.; Garcia-Parejo, P.; Levy, D. Preventing UV-light damage of light sensitive materials using a highly protective UV-absorbing coating. Chem. Soc. Rev. 2007, 36, 1270–1281. [Google Scholar] [CrossRef]
  39. Andrady, A.L.; Hamid, H.S.; Torikai, A. Effects of climate change and UV-B on materials. Photochem. Photobiol. Sci. 2003, 2, 68–72. [Google Scholar] [CrossRef]
  40. Andrady, A.L.; Hamid, H.; Torikai, A. Effects of solar UV and climate change on materials. Photochem. Photobiol. Sci. 2011, 10, 292–300. [Google Scholar] [CrossRef]
  41. Wang, C.; Astruc, D. Nanogold plasmonic photocatalysis for organic synthesis and clean energy conversion. Chem. Soc. Rev. 2014, 43, 7188–7216. [Google Scholar] [CrossRef] [Green Version]
  42. Rycenga, M.; Cobley, C.M.; Zeng, J.; Li, W.; Moran, C.H.; Zhang, Q.; Qin, D.; Xia, Y. Controlling the synthesis and assembly of silver nanostructures for plasmonic applications. Chem. Rev. 2011, 111, 3669–3712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Sherry, L.J.; Chang, S.-H.; Schatz, G.C.; Van Duyne, R.P.; Wiley, B.J.; Xia, Y. Localized surface plasmon resonance spectroscopy of single silver nanocubes. Nano Lett. 2005, 5, 2034–2038. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, Y.; He, S.; Guo, W.; Hu, Y.; Huang, J.; Mulcahy, J.R.; Wei, W.D. Surface-plasmon-driven hot electron photochemistry. Chem. Rev. 2018, 118, 2927–2954. [Google Scholar] [CrossRef] [PubMed]
  45. Kale, M.J.; Avanesian, T.; Christopher, P. Direct photocatalysis by plasmonic nanostructures. ACS Catal. 2014, 4, 116–128. [Google Scholar] [CrossRef]
  46. Chen, H.; Shao, L.; Li, Q.; Wang, J. Gold nanorods and their plasmonic properties. Chem. Soc. Rev. 2013, 42, 2679–2724. [Google Scholar] [CrossRef] [PubMed]
  47. Nehl, C.L.; Hafner, J.H. Shape-dependent plasmon resonances of gold nanoparticles. J. Mater. Chem. 2008, 18, 2415–2419. [Google Scholar] [CrossRef] [Green Version]
  48. Farokhnezhad, M.; Esmaeilzadeh, M.; Nourian, M.; Jalaeikhoo, H.; Rajaeinejad, M.; Iravani, S.; Majidzadeh-A, K. Silica–gold nanoshell@graphene: A novel class of plasmonic nanoagents for photothermal cancer therapy. J. Phys. D Appl. Phys. 2020, 53, 405401. [Google Scholar] [CrossRef]
  49. Oldenburg, S.J.; Averitt, R.D.; Westcott, S.L.; Halas, N.J. Nanoengineering of optical resonances. Chem. Phys. Lett. 1998, 288, 243–247. [Google Scholar] [CrossRef]
  50. Chen, J.; Saeki, F.; Wiley, B.J.; Cang, H.; Cobb, M.J.; Li, Z.-Y.; Au, L.; Zhang, H.; Kimmey, M.B.; Li, X.; et al. Gold nanocages: Bioconjugation and their potential use as optical imaging contrast agents. Nano Lett. 2005, 5, 473–477. [Google Scholar] [CrossRef]
  51. Brongersma, M.L.; Halas, N.J.; Nordlander, P. Plasmon-induced hot carrier science and technology. Nat. Nanotechnol. 2015, 10, 25–34. [Google Scholar] [CrossRef]
  52. Clavero, C. Plasmon-induced hot-electron generation at nanoparticle/metal-oxide interfaces for photovoltaic and photocatalytic devices. Nat. Photonics 2014, 8, 95–103. [Google Scholar] [CrossRef]
  53. Vitousek, P.M.; Mooney, H.A.; Lubchenco, J.; Melillo, J.M. Human domination of Earth’s ecosystems. Science 1997, 277, 494–499. [Google Scholar] [CrossRef] [Green Version]
  54. Lewis, N.S.; Nocera, D.G. Powering the planet: Chemical challenges in solar energy utilization. Proc. Natl. Acad. Sci. USA 2006, 103, 15729–15735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Wang, W.-H.; Himeda, Y.; Muckerman, J.T.; Manbeck, G.F.; Fujita, E. CO2 hydrogenation to formate and methanol as an alternative to photo- and electrochemical CO2 reduction. Chem. Rev. 2015, 115, 12936–12973. [Google Scholar] [CrossRef]
  56. Nielsen, C.J.; Herrmann, H.; Weller, C. Atmospheric chemistry and environmental impact of the use of amines in carbon capture and storage (CCS). Chem. Soc. Rev. 2012, 41, 6684–6704. [Google Scholar] [CrossRef]
  57. Wu, J.; Huang, Y.; Ye, W.; Li, Y. CO2 reduction: From the electrochemical to photochemical approach. Adv. Sci. 2017, 4, 1700194. [Google Scholar] [CrossRef]
  58. Dahl, S.; Chorkendorff, I. Towards practical implementation. Nat. Mater. 2012, 11, 100–101. [Google Scholar] [CrossRef]
  59. Schneider, J.; Jia, H.; Muckerman, J.T.; Fujita, E. Thermodynamics and kinetics of CO2, CO, and H+ binding to the metal centre of CO2 reduction catalysts. Chem. Soc. Rev. 2012, 41, 2036–2051. [Google Scholar] [CrossRef]
  60. Hu, B.; Guild, C.; Suib, S.L. Thermal, electrochemical, and photochemical conversion of CO2 to fuels and value-added products. J. CO2 Util. 2013, 1, 18–27. [Google Scholar] [CrossRef]
  61. Li, X.; Wen, J.; Low, J.; Fang, Y.; Yu, J. Design and fabrication of semiconductor photocatalyst for photocatalytic reduction of CO2 to solar fuel. Sci. China Mater. 2014, 57, 70–100. [Google Scholar] [CrossRef] [Green Version]
  62. Gao, D.; Arán-Ais, R.M.; Jeon, H.S.; Cuenya, B.R. Rational catalyst and electrolyte design for CO2 electroreduction towards multicarbon products. Nat. Catal. 2019, 2, 198–210. [Google Scholar] [CrossRef]
  63. Hepburn, C.; Adlen, E.; Beddington, J.; Carter, E.A.; Fuss, S.; Mac Dowell, N.; Minx, J.C.; Smith, P.; Williams, C.K. The technological and economic prospects for CO2 utilization and removal. Nature 2019, 575, 87–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Morris, A.J.; Meyer, G.J.; Fujita, E. Molecular approaches to the photocatalytic reduction of carbon dioxide for solar fuels. Acc. Chem. Res. 2009, 42, 1983–1994. [Google Scholar] [CrossRef] [PubMed]
  65. Kiesgen de_Richter, R.; Ming, T.; Caillol, S. Fighting global warming by photocatalytic reduction of CO2 using giant photocatalytic reactors. Renew. Sustain. Energy Rev. 2013, 19, 82–106. [Google Scholar] [CrossRef]
  66. Marszewski, M.; Cao, S.; Yu, J.; Jaroniec, M. Semiconductor-based photocatalytic CO2 conversion. Mater. Horiz. 2015, 2, 261–278. [Google Scholar] [CrossRef]
  67. Razzaq, A.; In, S.-I. TiO2 based nanostructures for photocatalytic CO2 conversion to valuable chemicals. Micromachines 2019, 10, 326. [Google Scholar] [CrossRef] [Green Version]
  68. Nguyen, T.P.; Nguyen, D.L.T.; Nguyen, V.-H.; Le, T.-H.; Vo, D.-V.N.; Trinh, Q.T.; Bae, S.-R.; Chae, S.Y.; Kim, S.Y.; Le, Q.V. Recent advances in TiO2-based photocatalysts for reduction of CO2 to fuels. Nanomaterials 2020, 10, 337. [Google Scholar] [CrossRef] [Green Version]
  69. Kumar, D.; Lee, A.; Lee, T.; Lim, M.; Lim, D.-K. Ultrafast and efficient transport of hot plasmonic electrons by graphene for Pt Free, highly efficient visible-light responsive photocatalyst. Nano Lett. 2016, 16, 1760–1767. [Google Scholar] [CrossRef]
  70. Yu, S.; Wilson, A.J.; Kumari, G.; Zhang, X.; Jain, P.K. Opportunities and challenges of solar-energy-driven carbon dioxide to fuel conversion with plasmonic catalysts. ACS Energy Lett. 2017, 2, 2058–2070. [Google Scholar] [CrossRef]
  71. Liu, E.; Hu, Y.; Li, H.; Tang, C.; Hu, X.; Fan, J.; Chen, Y.; Bian, J. Photoconversion of CO2 to methanol over plasmonic Ag/TiO2 nano-wire films enhanced by overlapped visible-light-harvesting nanostructures. Ceram. Int. 2015, 41, 1049–1057. [Google Scholar] [CrossRef]
  72. Louis, C.; Pluchery, O. Gold Nanoparticles for Physics, Chemistry, and Biology; Imperial College Press: London, UK, 2012. [Google Scholar]
  73. Sarina, S.; Waclawik, E.R.; Zhu, H. Photocatalysis on supported gold and silver nanoparticles under ultraviolet and visible light irradiation. Green Chem. 2013, 15, 1814–1833. [Google Scholar] [CrossRef]
  74. Cao, L.; Sahu, S.; Anilkumar, P.; Bunker, C.E.; Xu, J.; Fernando, K.A.S.; Wang, P.; Guliants, E.A.; Tackett, K.N.; Sun, Y.-P. Carbon nanoparticles as visible-light photocatalysts for efficient CO2 conversion and beyond. J. Am. Chem. Soc. 2011, 133, 4754–4757. [Google Scholar] [CrossRef] [PubMed]
  75. Hoffmann, M.R.; Moss, J.A.; Baum, M.M. Artificial photosynthesis: Semiconductor photocatalytic fixation of CO2 to afford higher organic compounds. Dalton Trans. 2011, 40, 5151–5158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Tu, W.; Zhou, Y.; Li, H.; Li, P.; Zou, Z. Au@TiO2 yolk–shell hollow spheres for plasmon-induced photocatalytic reduction of CO2 to solar fuel via a local electromagnetic field. Nanoscale 2015, 7, 14232–14236. [Google Scholar] [CrossRef] [PubMed]
  77. Yu, S.; Wilson, A.J.; Heo, J.; Jain, P.K. Plasmonic control of multi-electron transfer and C-C coupling in visible-light-driven CO2 reduction on Au nanoparticles. Nano Lett. 2018, 18, 2189–2194. [Google Scholar] [CrossRef]
  78. Zhao, J.; Liu, B.; Meng, L.; He, S.; Yuan, R.; Hou, Y.; Ding, Z.; Lin, H.; Zhang, Z.; Wang, X.; et al. Plasmonic control of solar-driven CO2 conversion at the metal/ZnO interfaces. Appl. Catal. B 2019, 256, 117823. [Google Scholar] [CrossRef]
  79. Becerra, J.; Nguyen, D.-T.; Gopalakrishnan, V.-N.; Do, T.-O. Plasmonic Au nanoparticles incorporated in the zeolitic imidazolate framework (ZIF-67) for the efficient sunlight-driven photoreduction of CO2. ACS Appl. Energy Mater. 2020, 3, 7659–7665. [Google Scholar] [CrossRef]
  80. Linic, S.; Christopher, P.; Ingram, D.B. Plasmonic-metal nanostructures for efficient conversion of solar to chemical energy. Nat. Mater. 2011, 10, 911–921. [Google Scholar] [CrossRef]
  81. Cai, B.; Wang, J.; Han, D.; Gan, S.; Zhang, Q.; Wu, Z.; Niu, L. Ternary alloyed AgClxBr1−x nanocrystals: Facile modulation of electronic structures toward advanced photocatalytic performance. Nanoscale 2013, 5, 10989–10995. [Google Scholar] [CrossRef]
  82. Cai, B.; Wang, J.; Gan, S.; Han, D.; Wu, Z.; Niu, L. A distinctive red Ag/AgCl photocatalyst with efficient photocatalytic oxidative and reductive activities. J. Mater. Chem. A 2014, 2, 5280–5286. [Google Scholar] [CrossRef]
  83. Li, N.; Jiang, R.; Li, Y.; Zhou, J.; Ma, Q.; Shen, S.; Liu, M. Plasma-assisted photocatalysis of CH4 and CO2 into ethylene. ACS Sustain. Chem. Eng. 2019, 7, 11455–11463. [Google Scholar] [CrossRef]
  84. Varghese, O.K.; Paulose, M.; LaTempa, T.J.; Grimes, C.A. High-rate solar photocatalytic conversion of CO2 and water vapor to hydrocarbon fuels. Nano Lett. 2009, 9, 731–737. [Google Scholar] [CrossRef] [PubMed]
  85. Tseng, I.-H.; Wu, J.C.-S. Chemical states of metal-loaded titania in the photoreduction of CO2. Catal. Today 2004, 97, 113–119. [Google Scholar] [CrossRef]
  86. Lee, S.; Kim, D.; Lee, J. Electrocatalytic production of C3-C4 compounds by conversion of CO2 on a chloride-induced Bi-phasic Cu2O-Cu catalyst. Angew. Chem. 2015, 127, 14914–14918. [Google Scholar] [CrossRef]
  87. Rawat, K.S.; Mahata, A.; Pathak, B. Thermochemical and electrochemical CO2 reduction on octahedral Cu nanocluster: Role of solvent towards product selectivity. J. Catal. 2017, 349, 118–127. [Google Scholar] [CrossRef]
  88. Shown, I.; Hsu, H.-C.; Chang, Y.-C.; Lin, C.-H.; Roy, P.K.; Ganguly, A.; Wang, C.-H.; Chang, J.-K.; Wu, C.-I.; Chen, L.-C.; et al. Highly efficient visible light photocatalytic reduction of CO2 to hydrocarbon fuels by Cu nanoparticle decorated graphene oxide. Nano Lett. 2014, 14, 6097–6103. [Google Scholar] [CrossRef]
  89. Park, H.; Ou, H.-H.; Colussi, A.J.; Hoffmann, M.R. Artificial photosynthesis of C1–C3 hydrocarbons from water and CO2 on titanate nanotubes decorated with nanoparticle elemental copper and CdS quantum dots. J. Phys. Chem. A 2015, 119, 4658–4666. [Google Scholar] [CrossRef]
  90. Chen, Y.; Wang, M.; Ma, Y.; Li, Y.; Cai, J.; Li, Z. Coupling photocatalytic CO2 reduction with benzyl alcohol oxidation to produce benzyl acetate over Cu2O/Cu. Catal. Sci. Technol. 2018, 8, 2218–2223. [Google Scholar] [CrossRef]
  91. Maluenda, I.; Navarro, O. Recent developments in the Suzuki–Miyaura reaction: 2010–2014. Molecules 2015, 20, 7528–7557. [Google Scholar] [CrossRef]
  92. Teranishi, T.; Eguchi, M.; Kanehara, M.; Gwo, S. Controlled localized surface plasmon resonance wavelength for conductive nanoparticles over the ultraviolet to near-infrared region. J. Mater. Chem. 2011, 21, 10238–10242. [Google Scholar] [CrossRef]
  93. Zielińska-Jurek, A. Progress, challenge, and perspective of bimetallic TiO2-based photocatalysts. J. Nanomater. 2014, 2014, 208920. [Google Scholar] [CrossRef] [Green Version]
  94. Wali, L.A.; Alwan, A.M.; Dheyab, A.B.; Hashim, D.A. Excellent fabrication of Pd-Ag NPs/PSi photocatalyst based on bimetallic nanoparticles for improving methylene blue photocatalytic degradation. Optik 2019, 179, 708–717. [Google Scholar] [CrossRef]
  95. Lee, S.; Wy, Y.; Lee, Y.W.; Ham, K.; Han, S.W. Core–shell nanoparticle clusters enable synergistic integration of plasmonic and catalytic functions in a single platform. Small 2017, 13, 1701633. [Google Scholar] [CrossRef]
  96. Furube, A.; Hashimoto, S. Insight into plasmonic hot-electron transfer and plasmon molecular drive: New dimensions in energy conversion and nanofabrication. NPG Asia Mater. 2017, 9, e454. [Google Scholar] [CrossRef]
  97. Shan, H.; Yu, Y.; Wang, X.; Luo, Y.; Zu, S.; Du, B.; Han, T.; Li, B.; Li, Y.; Wu, J.; et al. Direct observation of ultrafast plasmonic hot electron transfer in the strong coupling regime. Light Sci. Appl. 2019, 8, 9. [Google Scholar] [CrossRef] [Green Version]
  98. Huang, X.; Li, Y.; Chen, Y.; Zhou, H.; Duan, X.; Huang, Y. Plasmonic and catalytic AuPd nanowheels for the efficient conversion of light Into chemical energy. Angew. Chem. Int. Ed. 2013, 52, 6063–6067. [Google Scholar] [CrossRef]
  99. Gangishetty, M.K.; Fontes, A.M.; Malta, M.; Kelly, T.L.; Scott, R.W.J. Improving the rates of Pd-catalyzed reactions by exciting the surface plasmons of AuPd bimetallic nanotriangles. RSC Adv. 2017, 7, 40218–40226. [Google Scholar] [CrossRef] [Green Version]
  100. Sarina, S.; Zhu, H.; Jaatinen, E.; Xiao, Q.; Liu, H.; Jia, J.; Chen, C.; Zhao, J. Enhancing catalytic performance of palladium in gold and palladium alloy nanoparticles for organic synthesis reactions through visible light irradiation at ambient temperatures. J. Am. Chem. Soc. 2013, 135, 5793–5801. [Google Scholar] [CrossRef]
  101. Xiao, Q.; Sarina, S.; Jaatinen, E.; Jia, J.; Arnold, D.P.; Liu, H.; Zhu, H. Efficient photocatalytic Suzuki cross-coupling reactions on Au-Pd alloy nanoparticles under visible light irradiation. Green Chem. 2014, 16, 4272–4285. [Google Scholar] [CrossRef]
  102. Xiao, Q.; Sarina, S.; Bo, A.; Jia, J.; Liu, H.; Arnold, D.P.; Huang, Y.; Wu, H.; Zhu, H. Visible light-driven cross-coupling reactions at lower temperatures using a photocatalyst of palladium and gold alloy nanoparticles. ACS Catal. 2014, 4, 1725–1734. [Google Scholar] [CrossRef]
  103. Sarina, S.; Jaatinen, E.; Xiao, Q.; Huang, Y.M.; Christopher, P.; Zhao, J.C.; Zhu, H.Y. Photon energy threshold in direct photocatalysis with metal nanoparticles: Key evidence from the action spectrum of the reaction. J. Phys. Chem. Lett. 2017, 8, 2526–2534. [Google Scholar] [CrossRef]
  104. Zhang, S.; Chang, C.; Huang, Z.; Ma, Y.; Gao, W.; Li, J.; Qu, Y. Visible-light-activated Suzuki–Miyaura coupling reactions of aryl chlorides over the multifunctional Pd/Au/porous nanorods of CeO2 catalysts. ACS Catal. 2015, 5, 6481–6488. [Google Scholar] [CrossRef]
  105. Movahed, S.K.; Miraghaee, S.; Dabiri, M. AuPd alloy nanoparticles decorated graphitic carbon nitride as an excellent photocatalyst for the visible-light-enhanced Suzuki–Miyaura cross-coupling reaction. J. Alloys Compd. 2020, 819, 152994. [Google Scholar] [CrossRef]
  106. Subudhi, S.; Mansingh, S.; Tripathy, S.P.; Mohanty, A.; Mohapatra, P.; Rath, D.; Parida, K. The fabrication of Au/Pd plasmonic alloys on UiO-66-NH2: An efficient visible light-induced photocatalyst towards the Suzuki–Miyaura coupling reaction under ambient conditions. Catal. Sci. Technol. 2019, 9, 6585–6597. [Google Scholar] [CrossRef]
  107. Kataria, M.; Kumar, M.; Singh, Z.; Bhalla, V. Gold nanoparticles immobilized polymeric PBI derivative: Productive, portable, and photocatalytic system for Heck coupling. ACS Sustain. Chem. Eng. 2018, 6, 8223–8229. [Google Scholar] [CrossRef]
  108. Nemygina, N.A.; Nikoshvili, L.Z.; Tiamina, I.Y.; Bykov, A.V.; Smirnov, I.S.; LaGrange, T.; Kaszkur, Z.; Matveeva, V.G.; Sulman, E.M.; Kiwi-Minsker, L. Au core–Pd shell bimetallic nanoparticlesi within hyper-cross-linked polystyrene for mechanistic study of Suzuki cross-coupling: Homogeneous or heterogeneous catalysis? Org. Process Res. Dev. 2018, 22, 1606–1613. [Google Scholar] [CrossRef]
  109. Han, D.; Bao, Z.; Xing, H.; Yang, Y.; Ren, Q.; Zhang, Z. Fabrication of plasmonic Au-Pd alloy nanoparticles for photocatalytic Suzuki–Miyaura reactions under ambient conditions. Nanoscale 2017, 9, 6026–6032. [Google Scholar] [CrossRef]
  110. Tyagi, A.; Yamamoto, A.; Yoshida, H. Photocatalytic Ullmann coupling of aryl halides by a novel blended catalyst consisting of a TiO2 photocatalyst and an Al2O3 supported Pd-Au bimetallic catalyst. Catal. Sci. Technol. 2018, 8, 6196–6203. [Google Scholar] [CrossRef] [Green Version]
  111. Tyagi, A.; Yamamoto, A.; Yoshida, H. Novel blended catalysts consisting of a TiO2 photocatalyst and an Al2O3 supported Pd-Au bimetallic catalyst for direct dehydrogenative cross-coupling between arenes and tetrahydrofuran. RSC Adv. 2018, 8, 24021–24028. [Google Scholar] [CrossRef] [Green Version]
  112. Rohani, S.; Ziarati, A.; Ziarani, G.M.; Badiei, A.; Burgi, T. Engineering of highly active Au/Pd supported on hydrogenated urchin-like yolk@shell TiO2 for visible light photocatalytic Suzuki coupling. Catal. Sci. Technol. 2019, 9, 3820–3827. [Google Scholar] [CrossRef] [Green Version]
  113. Verma, P.; Kuwahara, Y.; Mori, K.; Yamashita, H. Pd/Ag and Pd/Au bimetallic nanocatalysts on mesoporous silica for plasmon-mediated enhanced catalytic activity under visible light irradiation. J. Mater. Chem. A 2016, 4, 10142–10150. [Google Scholar] [CrossRef]
  114. Li, X.-H.; Baar, M.; Blechert, S.; Antonietti, M. Facilitating room-temperature Suzuki coupling reaction with light: Mott-Schottky photocatalyst for C-C coupling. Sci. Rep. 2013, 3, 1743. [Google Scholar] [CrossRef]
  115. Shin, H.H.; Kang, E.; Park, H.; Han, T.; Lee, C.-H.; Lim, D.-K. Pd-nanodot decorated MoS2 nanosheets as a highly efficient photocatalyst for the visible-light-induced Suzuki–Miyaura coupling reaction. J. Mater. Chem. A 2017, 5, 24965–24971. [Google Scholar] [CrossRef]
  116. Raza, F.; Yim, D.; Park, J.H.; Kim, H.-I.; Jeon, S.-J.; Kim, J.-H. Structuring Pd nanoparticles on 2H-WS2 nanosheets induces excellent photocatalytic activity for cross-coupling reactions under visible light. J. Am. Chem. Soc. 2017, 139, 14767–14774. [Google Scholar] [CrossRef]
  117. Chen, J.-H.; Jang, C.; Xiao, S.; Ishigami, M.; Fuhrer, M.S. Intrinsic and extrinsic performance limits of graphene devices on SiO2. Nat. Nanotechnol. 2008, 3, 206–209. [Google Scholar] [CrossRef]
  118. Xiang, Q.; Yu, J.; Jaroniec, M. Graphene-based semiconductor photocatalysts. Chem. Soc. Rev. 2012, 41, 782–796. [Google Scholar] [CrossRef]
  119. Kang, E.; Shin, H.H.; Lim, D.-K. Interface-controlled Pd nanodot-Au nanoparticle colloids for efficient visible-light-induced photocatalytic Suzuki–Miyaura coupling reaction. Catalysts 2018, 8, 463. [Google Scholar] [CrossRef] [Green Version]
  120. Sahoo, M.; Mansingh, S.; Subudhi, S.; Mohapatra, P.; Parida, K. A plasmonic AuPd bimetallic nanoalloy decorated over a GO/LDH hybrid nanocomposite via a green synthesis route for robust Suzuki coupling reactions: A paradigm shift towards a sustainable future. Catal. Sci. Technol. 2019, 9, 4678–4692. [Google Scholar] [CrossRef]
  121. Shin, H.H.; Han, T.; Yang, W.; Lim, D.-K. Pd2+-conjugated reduced graphene oxide as a highly efficient visible-light-induced photocatalytic Suzuki–Miyaura coupling. Carbon 2019, 143, 362–370. [Google Scholar] [CrossRef]
  122. Wang, L.; Xu, X.; Cheng, Q.; Dou, S.X.; Du, Y. Near-infrared-driven photocatalysts: Design, construction, and applications. Small 2019, 1904107. [Google Scholar] [CrossRef]
  123. Murphy, C.J.; Sau, T.K.; Gole, A.M.; Orendorff, C.J.; Gao, J.; Gou, L.; Hunyadi, S.E.; Li, T. Anisotropic metal nanoparticles: Synthesis, assembly, and optical applications. J. Phys. Chem. B 2005, 109, 13857–13870. [Google Scholar] [CrossRef]
  124. Khlebtsov, B.N.; Khlebtsov, N.G. Multipole plasmons in metal nanorods: Scaling properties and dependence on particle size, shape, orientation, and dielectric environment. J. Phys. Chem. C 2007, 111, 11516–11527. [Google Scholar] [CrossRef]
  125. Encina, E.R.; Coronado, E.A. Resonance conditions for multipole plasmon excitations in noble metal nanorods. J. Phys. Chem. C 2007, 111, 16796–16801. [Google Scholar] [CrossRef]
  126. Wang, J.-H.; Chen, M.; Luo, Z.-J.; Ma, L.; Zhang, Y.-F.; Chen, K.; Zhou, L.; Wang, Q.-Q. Ceria-coated gold nanorods for plasmon-enhanced near-infrared photocatalytic and photoelectrochemical performances. J. Phys. Chem. C 2016, 120, 14805–14812. [Google Scholar] [CrossRef]
  127. Wang, F.; Li, C.; Chen, H.; Jiang, R.; Sun, L.-D.; Li, Q.; Wang, J.; Yu, J.C.; Yan, C.-H. Plasmonic harvesting of light energy for Suzuki coupling reactions. J. Am. Chem. Soc. 2013, 135, 5588–5601. [Google Scholar] [CrossRef]
  128. Wen, M.; Takakura, S.; Fuku, K.; Mori, K.; Yamashita, H. Enhancement of Pd-catalyzed Suzuki–Miyaura coupling reaction assisted by localized surface plasmon resonance of Au nanorods. Catal. Today 2015, 242, 381–385. [Google Scholar] [CrossRef]
  129. Guo, J.; Zhang, Y.; Shi, L.; Zhu, Y.; Mideksa, M.F.; Hou, K.; Zhao, W.; Wang, D.; Zhao, M.; Zhang, X.; et al. Boosting hot electrons in hetero-superstructures for plasmon-enhanced catalysis. J. Am. Chem. Soc. 2017, 139, 17964–17972. [Google Scholar] [CrossRef]
  130. Verkaaik, M.; Grote, R.; Meulendijks, N.; Sastre, F.; Weckhuysen, B.M.; Buskens, P. Suzuki-Miyaura cross-coupling using plasmonic Pd-decorated Au nanorods as catalyst: A study on the contribution of laser illumination. ChemCatChem 2019, 11, 4974–4980. [Google Scholar] [CrossRef]
  131. Yoshii, T.; Kuwahara, Y.; Mori, K.; Yamashita, H. Design of Pd–graphene–Au nanorod nanocomposite catalyst for boosting Suzuki–Miyaura coupling reaction by assistance of surface plasmon resonance. J. Phys. Chem. C 2019, 123, 24575–24583. [Google Scholar] [CrossRef]
  132. Gao, S.; Shang, N.; Feng, C.; Wang, C.; Wang, Z. Graphene oxide–palladium modified Ag-AgBr: A visible-light-responsive photocatalyst for the Suzuki coupling reaction. RSC Adv. 2014, 4, 39242–39247. [Google Scholar] [CrossRef]
  133. Verma, P.; Kuwahara, Y.; Mori, K.; Yamashita, H. Synthesis and characterization of a Pd/Ag bimetallic nanocatalyst on SBA-15 mesoporous silica as a plasmonic catalyst. J. Mater. Chem. A 2015, 3, 18889–18897. [Google Scholar] [CrossRef]
  134. Sharma, K.; Kumar, M.; Bhalla, V. Aggregates of the pentacenequinone derivative as reactors for the preparation of Ag@Cu2O core-shell NPs: An active photocatalyst for Suzuki and Suzuki type coupling reactions. Chem. Commun. 2015, 51, 12529–12532. [Google Scholar] [CrossRef] [PubMed]
  135. Chopra, R.; Kumar, M.; Bhalla, V. Development of a supramolecular ensemble of an AIEE active hexaphenylbenzene derivative and Ag@Cu2O Ccore-shell NPs: An efficient photocatalytic system for C-H activation. Chem. Commun. 2016, 52, 10179–10182. [Google Scholar] [CrossRef]
  136. Chen, Y.; Feng, L. Silver nanoparticles doped TiO2 catalyzed Suzuki coupling of bromoaryl with phenylboronic acid under visible light. J. Photochem. Photobiol. B 2020, 205, 111807. [Google Scholar] [CrossRef] [PubMed]
  137. Chan, G.H.; Zhao, J.; Hicks, E.M.; Schatz, G.C.; Van Duyne, R.P. Plasmonic properties of copper nanoparticles fabricated by nanosphere lithography. Nano Lett. 2007, 7, 1947–1952. [Google Scholar] [CrossRef]
  138. Liu, P.; Wang, H.; Li, X.; Rui, M.; Zeng, H. Localized surface plasmon resonance of Cu nanoparticles by laser ablation in liquid media. RSC Adv. 2015, 5, 79738–79745. [Google Scholar] [CrossRef]
  139. Gezgin, S.Y.; Kepceoğlu, A.; Kılıç, H.Ş. An experimental investigation of localised surface plasmon resonance (LSPR) for Cu nanoparticles depending as a function of laser pulse number in pulsed laser deposition. AIP Conf. Proc. 2017, 1815, 030020. [Google Scholar]
  140. Kim, J.H.; Ehrman, S.H.; Germer, T.A. Influence of particle oxide coating on light scattering by submicron metal particles on silicon wafers. Appl. Phys. Lett. 2004, 84, 1278–1280. [Google Scholar] [CrossRef]
  141. DeSario, P.A.; Pietron, J.J.; Brintlinger, T.H.; McEntee, M.; Parker, J.F.; Baturina, O.; Stroud, R.M.; Rolison, D.R. Oxidation-stable plasmonic copper nanoparticles in photocatalytic TiO2 nanoarchitectures. Nanoscale 2017, 9, 11720–11729. [Google Scholar] [CrossRef] [Green Version]
  142. Huang, Y.; Liu, Z.; Gao, G.; Xiao, G.; Du, A.; Bottle, S.; Sarina, S.; Zhu, H. Stable copper nanoparticle photocatalysts for selective epoxidation of alkenes with visible light. ACS Catal. 2017, 7, 4975–4985. [Google Scholar] [CrossRef]
  143. Zhang, P.; Song, T.; Wang, T.; Zeng, H. Plasmonic Cu nanoparticle on reduced graphene oxide nanosheet support: An efficient photocatalyst for improvement of near-infrared photocatalytic H2 evolution. Appl. Catal. B 2018, 225, 172–179. [Google Scholar] [CrossRef]
  144. Wang, B.; Wang, Y.; Li, J.; Guo, X.; Bai, G.; Tong, X.; Jin, G.; Guo, X. Photocatalytic Sonogashira reaction over silicon carbide supported Pd-Cu alloy nanoparticles under visible light irradiation. Catal. Sci. Technol. 2018, 8, 3357–3362. [Google Scholar] [CrossRef]
  145. Sun, D.; Xu, M.; Jiang, Y.; Long, J.; Li, Z. Small-sized bimetallic CuPd nanoclusters encapsulated inside cavity of NH2-UiO-66(Zr) with superior performance for light-induced Suzuki coupling reaction. Small Methods 2018, 2, 1800164. [Google Scholar] [CrossRef]
  146. Guo, X.; Hao, C.; Jin, G.; Zhu, H.-Y.; Guo, X.-Y. Copper nanoparticles on graphene support: An efficient photocatalyst for coupling of nitroaromatics in visible light. Angew. Chem. Int. Ed. 2014, 53, 1973–1977. [Google Scholar] [CrossRef] [PubMed]
  147. Cui, Y.-L.; Guo, X.-N.; Wang, Y.-Y.; Guo, X.-Y. Visible-light-driven photocatalytic N-arylation of imidazole derivatives and arylboronic acids on Cu/graphene catalyst. Sci. Rep. 2015, 5, 12005. [Google Scholar] [CrossRef] [Green Version]
  148. Kaur, S.; Kumar, M.; Bhalla, V. Supramolecular ensemble of PBI derivative and copper nanoparticles: A light harvesting antenna for photocatalytic C(sp2)–H functionalization. Green Chem. 2016, 18, 5870–5883. [Google Scholar] [CrossRef]
  149. Li, M.; Xing, X.; Ma, Z.; Lv, J.; Fu, P.; Li, Z. Synthesis of composition tunable and (111)-faceted Cu/Cu2O nanoparticles toward photocatalytic, ligand-free, and solvent-free C-N Ullmann coupling reactions. ACS Sustain. Chem. Eng. 2018, 6, 5495–5503. [Google Scholar] [CrossRef]
  150. Trinh, T.T.; Sato, R.; Sakamoto, M.; Fujiyoshi, Y.; Haruta, M.; Kurata, H.; Teranishi, T. Visible to near-infrared plasmon-enhanced catalytic activity of Pd hexagonal nanoplates for the Suzuki coupling reaction. Nanoscale 2015, 7, 12435–12444. [Google Scholar] [CrossRef]
  151. Cui, J.; Li, Y.; Liu, L.; Chen, L.; Xu, J.; Ma, J.; Fang, G.; Zhu, E.; Wu, H.; Zhao, L.; et al. Near-infrared plasmonic-enhanced solar energy harvest for highly efficient photocatalytic reactions. Nano Lett. 2015, 15, 6295–6301. [Google Scholar] [CrossRef]
  152. Lou, Z.; Gu, Q.; Liao, Y.; Yu, S.; Xue, C. Promoting Pd-catalyzed Suzuki coupling reactions through near-infrared plasmon excitation of WO3−x nanowires. Appl. Catal. B 2016, 184, 258–263. [Google Scholar] [CrossRef]
  153. Sarhid, I.; Abdellah, I.; Martini, C.; Huc, V.; Dragoe, D.; Beaunier, P.; Lampre, I.; Remita, H. Plasmonic catalysis for the Suzuki–Miyaura cross-coupling reaction using palladium nanoflowers. New J. Chem. 2019, 43, 4349–4355. [Google Scholar] [CrossRef]
  154. Koohgard, M.; Hosseini-Sarvari, M. Enhancement of Suzuki–Miyaura coupling reaction by photocatalytic palladium nanoparticles anchored to TiO2 under visible light irradiation. Catal. Commun. 2018, 111, 10–15. [Google Scholar] [CrossRef] [Green Version]
  155. Hosseini-Sarvari, M.; Bazyar, Z. Visible light driven photocatalytic cross-coupling reactions on nano Pd/ZnO photocatalyst at room-temperature. ChemistrySelect 2018, 3, 1898–1907. [Google Scholar] [CrossRef]
Figure 1. The carbon–carbon coupling reactions. (a) The general process of CO2 reduction in a photocatalytic system inspired by photosynthesis and (b) the summary of carbon–carbon cross-coupling in organic chemistry.
Figure 1. The carbon–carbon coupling reactions. (a) The general process of CO2 reduction in a photocatalytic system inspired by photosynthesis and (b) the summary of carbon–carbon cross-coupling in organic chemistry.
Catalysts 11 00155 g001
Figure 2. (a) The mechanism of CO2 photoreduction into CH4 and C2H6 with AuNPs (11.8 ± 2.3 nm), and the photocatalytic efficiency under (b) different wavelengths and (c,d) light intensities. Reproduced with permission from Reference [77]. Copyright 2018 American Chemical Society.
Figure 2. (a) The mechanism of CO2 photoreduction into CH4 and C2H6 with AuNPs (11.8 ± 2.3 nm), and the photocatalytic efficiency under (b) different wavelengths and (c,d) light intensities. Reproduced with permission from Reference [77]. Copyright 2018 American Chemical Society.
Catalysts 11 00155 g002
Figure 3. (a) Schematic illustration of the photocatalytic conversion of CO2 to C2H6 using Au/ZnO nanosheets, and the hydrocarbon production rates from CO2 with loading plasmonic metals of (b) Au, (c) Ag, and (d) Pd to ZnO. Reproduced with permission from Reference [78]. Copyright 2019 Elsevier.
Figure 3. (a) Schematic illustration of the photocatalytic conversion of CO2 to C2H6 using Au/ZnO nanosheets, and the hydrocarbon production rates from CO2 with loading plasmonic metals of (b) Au, (c) Ag, and (d) Pd to ZnO. Reproduced with permission from Reference [78]. Copyright 2019 Elsevier.
Catalysts 11 00155 g003
Figure 4. Photocatalytic activity of Aux@ZIF-67 for CO2 reduction to (a) CH3OH and (b) C2H5OH, and (c) the mechanism of CO2 photoreduction. Reproduced with permission from Reference [79]. Copyright 2020 American Chemical Society.
Figure 4. Photocatalytic activity of Aux@ZIF-67 for CO2 reduction to (a) CH3OH and (b) C2H5OH, and (c) the mechanism of CO2 photoreduction. Reproduced with permission from Reference [79]. Copyright 2020 American Chemical Society.
Catalysts 11 00155 g004
Figure 5. (a) The comparison of the photocatalytic efficiency for CO2 reduction to CH3OH and C2H5OH with red Ag/AgCl, and (b) a schematic illustration of physical photocatalytic mechanism. Reproduced with permission from Reference [82]. Copyright 2014 Royal Society of Chemistry.
Figure 5. (a) The comparison of the photocatalytic efficiency for CO2 reduction to CH3OH and C2H5OH with red Ag/AgCl, and (b) a schematic illustration of physical photocatalytic mechanism. Reproduced with permission from Reference [82]. Copyright 2014 Royal Society of Chemistry.
Catalysts 11 00155 g005
Figure 6. (a) Schematic illustration of Ag/TiO2 for photocatalysis of CH4 and CO2 into ethylene and the photocatalytic performance depending on (b) the light source, (c) primary photocatalyst, and (d) metal cocatalyst. Reproduced with permission from Reference [83]. Copyright 2019 American Chemical Society.
Figure 6. (a) Schematic illustration of Ag/TiO2 for photocatalysis of CH4 and CO2 into ethylene and the photocatalytic performance depending on (b) the light source, (c) primary photocatalyst, and (d) metal cocatalyst. Reproduced with permission from Reference [83]. Copyright 2019 American Chemical Society.
Catalysts 11 00155 g006
Figure 7. Au–Pd bimetallic alloys for photocatalytic Suzuki coupling. (a) AuPd nano-wheels. Reproduced with permission from Reference [98]. Copyright 2013 Wiley-VCH. (b) AuPd nanotriangles. Reproduced with permission from Reference [99]. Copyright 2017 Royal Society of Chemistry.
Figure 7. Au–Pd bimetallic alloys for photocatalytic Suzuki coupling. (a) AuPd nano-wheels. Reproduced with permission from Reference [98]. Copyright 2013 Wiley-VCH. (b) AuPd nanotriangles. Reproduced with permission from Reference [99]. Copyright 2017 Royal Society of Chemistry.
Catalysts 11 00155 g007
Figure 8. Au-Pd alloy nanoparticles (NPs) embedded on ZrO2 for visible-light-induced photocatalytic C–C cross-couplings. (a) Morphology, characterization, and photocatalytic efficiency of Suzuki coupling. Reproduced with permission from Reference [100]. Copyright 2013 American Chemical Society. Apparent activation energies of (b) Suzuki, (c) Sonogashira, and Hiyama couplings calculated for the photoreaction and the reaction in the dark. Reproduced with permission from References [101,102]. Copyright 2014 Royal Society of Chemistry. Copyright 2014 American Chemical Society. (d) The distribution of hot electron in the plasmonic metal or its alloy NPs with 400 nm, 530 nm irradiation, and their contribution to quantum efficiency. Reproduced with permission from Reference [103]. Copyright 2017 American Chemical Society.
Figure 8. Au-Pd alloy nanoparticles (NPs) embedded on ZrO2 for visible-light-induced photocatalytic C–C cross-couplings. (a) Morphology, characterization, and photocatalytic efficiency of Suzuki coupling. Reproduced with permission from Reference [100]. Copyright 2013 American Chemical Society. Apparent activation energies of (b) Suzuki, (c) Sonogashira, and Hiyama couplings calculated for the photoreaction and the reaction in the dark. Reproduced with permission from References [101,102]. Copyright 2014 Royal Society of Chemistry. Copyright 2014 American Chemical Society. (d) The distribution of hot electron in the plasmonic metal or its alloy NPs with 400 nm, 530 nm irradiation, and their contribution to quantum efficiency. Reproduced with permission from Reference [103]. Copyright 2017 American Chemical Society.
Catalysts 11 00155 g008
Figure 9. AuPd NPs decorated semiconductors for photocatalytic Suzuki coupling under visible light. (a) Pd/Au/CeO2. Reproduced with permission from Reference [104]. Copyright 2015 American Chemical Society. (b) AuPd/g-C3N4. Reproduced with permission from Reference [105]. Copyright 2020 Elsevier.
Figure 9. AuPd NPs decorated semiconductors for photocatalytic Suzuki coupling under visible light. (a) Pd/Au/CeO2. Reproduced with permission from Reference [104]. Copyright 2015 American Chemical Society. (b) AuPd/g-C3N4. Reproduced with permission from Reference [105]. Copyright 2020 Elsevier.
Catalysts 11 00155 g009
Figure 10. The HUY@S-TOH/AuPd for visible-light-driven photocatalytic Suzuki coupling. Reproduced with permission from Reference [112]. Copyright 2019 Royal Society of Chemistry.
Figure 10. The HUY@S-TOH/AuPd for visible-light-driven photocatalytic Suzuki coupling. Reproduced with permission from Reference [112]. Copyright 2019 Royal Society of Chemistry.
Catalysts 11 00155 g010
Figure 11. Pd-nanodot-decorated AuNPs with a graphene interface for visible-light-induced photocatalytic Suzuki–Miyaura coupling reaction. (a) The scheme of the Suzuki–Miyaura cross-coupling with Pd nanodot-decorated AuNPs and (b,c) the catalytic activities of Pd-cys-AuNPs, Pd-GO-AuNPs, and Pd-rGO-AuNPs for the reaction under various conditions. Reproduced with permission from Reference [119]. Copyright 2018 Creative Commons Attribution.
Figure 11. Pd-nanodot-decorated AuNPs with a graphene interface for visible-light-induced photocatalytic Suzuki–Miyaura coupling reaction. (a) The scheme of the Suzuki–Miyaura cross-coupling with Pd nanodot-decorated AuNPs and (b,c) the catalytic activities of Pd-cys-AuNPs, Pd-GO-AuNPs, and Pd-rGO-AuNPs for the reaction under various conditions. Reproduced with permission from Reference [119]. Copyright 2018 Creative Commons Attribution.
Catalysts 11 00155 g011
Figure 12. NIR-light-induced plasmonic photocatalysts for C–C cross-coupling. (a) Au–Pd nanostructures. Reproduced with permission from Reference [127]. Copyright 2013 American Chemical Society. (b) AuNR@Pd superstructures. Reproduced with permission from Reference [129]. Copyright 2017 American Chemical Society. (c) Pd-graphene-AuNR nanocomposites. Reproduced with permission from Reference [131]. Copyright 2019 American Chemical Society.
Figure 12. NIR-light-induced plasmonic photocatalysts for C–C cross-coupling. (a) Au–Pd nanostructures. Reproduced with permission from Reference [127]. Copyright 2013 American Chemical Society. (b) AuNR@Pd superstructures. Reproduced with permission from Reference [129]. Copyright 2017 American Chemical Society. (c) Pd-graphene-AuNR nanocomposites. Reproduced with permission from Reference [131]. Copyright 2019 American Chemical Society.
Catalysts 11 00155 g012
Figure 13. Pd/Ag bimetallic nano-catalysts on mesoporous silica for photocatalytic Suzuki coupling. (a) UV-vis spectra and wavelength dependence of the photocatalysts, (b) the catalytic activities of Pd/Ag/SBA-15 catalysts with different SBA-15, (c) the mechanism for the enhanced photocatalytic activity, and (d) the catalytic activities of Pd/metal/SBA-15. Reproduced with permission from References [113,133]. Copyright 2015 and 2016 Royal Society of Chemistry.
Figure 13. Pd/Ag bimetallic nano-catalysts on mesoporous silica for photocatalytic Suzuki coupling. (a) UV-vis spectra and wavelength dependence of the photocatalysts, (b) the catalytic activities of Pd/Ag/SBA-15 catalysts with different SBA-15, (c) the mechanism for the enhanced photocatalytic activity, and (d) the catalytic activities of Pd/metal/SBA-15. Reproduced with permission from References [113,133]. Copyright 2015 and 2016 Royal Society of Chemistry.
Catalysts 11 00155 g013
Figure 14. Ag@Cu2O core–shell nanoparticles (NPs) stabilized by a supramolecular ensemble for photocatalytic Suzuki coupling. Reproduced with permission from References [134,135]. Copyright 2015 and 2016 Royal Society of Chemistry.
Figure 14. Ag@Cu2O core–shell nanoparticles (NPs) stabilized by a supramolecular ensemble for photocatalytic Suzuki coupling. Reproduced with permission from References [134,135]. Copyright 2015 and 2016 Royal Society of Chemistry.
Catalysts 11 00155 g014
Figure 15. The schematic diagram of the proposed mechanism using CuPd bimetallic alloy for photocatalytic C–C cross-coupling. (a) PdCu/SiC for photocatalytic Sonogashira reaction. Reproduced with permission from Reference [144]. Copyright 2018 Royal Society of Chemistry. (b) CuPd@NH2-UiO-66(Zr) for photocatalytic Suzuki coupling. Reproduced with permission from Reference [145]. Copyright 2018 Wiley-VCH.
Figure 15. The schematic diagram of the proposed mechanism using CuPd bimetallic alloy for photocatalytic C–C cross-coupling. (a) PdCu/SiC for photocatalytic Sonogashira reaction. Reproduced with permission from Reference [144]. Copyright 2018 Royal Society of Chemistry. (b) CuPd@NH2-UiO-66(Zr) for photocatalytic Suzuki coupling. Reproduced with permission from Reference [145]. Copyright 2018 Wiley-VCH.
Catalysts 11 00155 g015
Figure 16. Plasmonic CuNPs with other support materials for photocatalysis. (a) Cu/graphene for photocatalytic C-N cross-coupling reaction. Reproduced with permission from Reference [147]. Copyright 2015 Nature Publishing Group. (b) Supramolecular ensemble 3:CuNPs for photocatalytic C-C cross-coupling by C(sp2)-H functionalization. Reproduced with permission from Reference [148]. Copyright 2016 Royal Society of Chemistry.
Figure 16. Plasmonic CuNPs with other support materials for photocatalysis. (a) Cu/graphene for photocatalytic C-N cross-coupling reaction. Reproduced with permission from Reference [147]. Copyright 2015 Nature Publishing Group. (b) Supramolecular ensemble 3:CuNPs for photocatalytic C-C cross-coupling by C(sp2)-H functionalization. Reproduced with permission from Reference [148]. Copyright 2016 Royal Society of Chemistry.
Catalysts 11 00155 g016
Table 1. Plasmonic hybrid nano-catalysts for various photocatalytic CO2 conversion into multi-carbon products. AuNP-, AgNP- and CuNP-assisted photocatalysts under visible light.
Table 1. Plasmonic hybrid nano-catalysts for various photocatalytic CO2 conversion into multi-carbon products. AuNP-, AgNP- and CuNP-assisted photocatalysts under visible light.
PhotocatalystSize and ShapeProductLight SourceReaction ConditionTimeYield
(μmoL g−1 h−1)
Ref.
Au@TiO2Spherical core: 45 nm
Spherical shell: 200–250 nm (50 nm thickness)
CH4, C2H6300 W Xe arc lampH2O10 h2.52 (CH4),
1.67 (C2H6)
[76]
AuNPs11.8 ± 2.3 nm spherical NPsCH4, C2H6300 W Xe lamp10% IPA10 h0.65 (CH4),
0.56 (C2H6) as TOF (NP−1h−1)
[77]
Au/ZnOAuNPs: 7 nm
ZnO sheets: >μm (1.6 nm thickness, 10–100 nm pores)
CH4, C2H6300 W Xe lamp
(>320 nm)
H2O1–6 h21.0 (CH4),
27.0 (C2H6)
[78]
Au20@ZIF-67AuNPs: 30–40 nm
ZIF-67: ~μm
CH3OH
C2H5OH
Solar simulator
(150 mW cm−2)
10 wt.% TEOA, 0.08 M NaHCO34 h1623 (CH3OH),
495 (C2H5OH)
[79]
AgCl0.75Br0.25Cubic nanocrystals
150–260 nm
CH3OH
C2H5OH
500 W Xe arc lamp
(>420 nm)
0.1 M NaHCO35 h36.2 (CH3OH),
72.4 (C2H5OH)
[81]
Ag/AgClCoaxial tri-cubic
500–600 nm
CH3OH
C2H5OH
500 W Xe arc lamp
(>420 nm)
0.1 M NaHCO35 h29.2 (CH3OH),
44.6 (C2H5OH)
[82]
Ag/TiO2AgNPs: 4 nm
TiO2: 25 nm
CO
C2H4
Xe lampQuartz cotton,
micro-
autoclave
2 h1149 (CO)
686 (C2H4)
[83]
Cu/GOCu (111) NPs: 5 nm
GO: >μm
CH3OH
CH3CHO
Halogen lamp (300 W)Continuous gas flow reactor2 h2.94 (CH3OH),
3.88 (CH3CHO)
[88]
CdS/(Cu-TNTs)Hexagonal CdSCH4
C2H6
C3H8
450 W Xe lamp
(>420 nm)
H2O5 h~28 (CH4),
~17 (C2H6),
~9 (C3H8)
μL g−1 h−1
[89]
Cu2O/CuIrregular porous structures
100 nm
Benzyl acetate300 W Xe lamp
(420–800 nm)
MeCN, benzyl alcohol20 h116.7[90]
Table 2. Plasmonic hybrid nano-catalysts for various photocatalytic C–C cross-coupling. AuNP-, AgNP-, CuNP-assisted photocatalysts under visible and NIR light.
Table 2. Plasmonic hybrid nano-catalysts for various photocatalytic C–C cross-coupling. AuNP-, AgNP-, CuNP-assisted photocatalysts under visible and NIR light.
PhotocatalystSize and ShapeReactionLight SourceReaction
Temp.
Solvent,
Base
TimeYieldTOF (h−1)Ref.
AuPd wheelsNano-wheels
290 nm (6 nm thickness)
SuzukiXe lamp50 °CEtOH/H2O
(9:1),
K2CO3
1 h65.8%-[98]
AuPd nanotrianglesNanotriangles
43 ± 4 nm
SuzukiGreen LED
(450–600 nm)
-EtOH/H2O
(1:6),
K2CO3
5 h>80%-[99]
Au–Pd alloy NPs/ZrO2Au-Pd NPs: <8 nmSuzuki500 W halogen lamp
(400–750 nm)
30 °CDMF/H2O
(3:1),
K2CO3
6 h96%14.5[100]
Au–Pd alloy NPs/ZrO2Au-Pd NPs: <7 nmSonogashira
or Stille
Halogen lamp
(400–750 nm)
45 °CH2O,
K3PO4
or NaOH
24 h80%
−81%
4.7
−4.8
[102]
Au–Pd alloy NPs/ZrO2Au-Pd NPs: <7 nmSuzukiHalogen lamp
(400–750 nm)
30 °CDMF/H2O
(3:1),
K2CO3
6 h96%14.5[101]
Pd/Au/CeO2AuNPs (111): 4.28 ± 1.05 nm
PdNPs (111): 5.14 ± 1.01 nm
CeO2 nanorods: ~5 nm (width), ~30 nm (length)
Suzuki150 W Xe lamp
(>400 nm)
25 °CDMF/H2O
(1:1),
K2CO3
5 h98.8%-[104]
Pd/Au/SBA-15Pd/Au NPs: 4.9 nm
SBA-15: >μm
SuzukiXe lampRoom Temp.EtOH
K2CO3
2 h70%-[113]
Au–Pd alloy NPs/TiO2Au-Pd NPs: 3 nmSuzuki5 W blue LED lamp25 °CEtOH/H2O
(1:1),
K2CO3
5 h98%-[109]
Pd-rGO-AuNPsPd nanodots: 2–3 nm
AuNPs: ~30 nm
SuzukiXe lamp
(400–800 nm)
25 °CEtOH/H2O
(1:1),
K2CO3
2 h54.5%-[119]
Au–Pd/HPSAuNPs core: ~4 nm
Pd shell: <1 nm
HPS: 15–50 nm
Suzuki300 W filament lamp60 °CEtOH/H2O
(5:1),
NaOH
3 h71.6%130.7[108]
TiO2 + PdAu/Al2O3-UllmannXe lamp
(≥350 nm)
Room Temp.CH3CN0.5 h2.2%-[110]
TiO2 + PdAu/Al2O3PdAu NPs: 3–4 nmDehydrogenative cross-couplingXe lamp
(≥350 nm)
Room Temp.-1 h15.2 μmol-[111]
Supramolecular Polymer 5:AuNPsSupramolecule: ~μm
AuNPs: <30 nm
Heck100 W tungsten filament blubRoom Temp.H2O,
K2CO3
1 h89%-[107]
GO/LDH
@AuPd
AuPd NPs: ~4.2 nmSuzuki300 W Xe lamp
(≥420 nm)
25 °CEtOH/H2O
(3:1),
K2CO3
2 h99.5%-[120]
HUY@S-TOH/AuPdAuNPs core: 5 nm
Pd shell: ~0.7 nm
HUY@S-TOH: ~3 μm
Suzuki300 W Xe lampRoom Temp.EtOH/H2O
(2:1),
K2CO3
0.5 h>99%7095[112]
Au/Pd@UiO-66-NH2Au/Pd NPs: 6.45 nm
UiO-66-NH2: ~50 nm
Suzuki300 W Xe lamp
(>420 nm)
25 °CEtOH/H2O
(1:1),
K2CO3
1 h>99%433[106]
AuPd/g-C3N4AuPd NPs: 5 nmSuzuki5 W Xe HID lamp25 °CEtOH/H2O
(1:1),
K3PO4
0.5 h99%7920[105]
Au–Pd nanostructuresAuNRs: 25
± 2 (diameter), 82 ± 6 nm (length)
PdNPs: 3–5 nm
SuzukiContinuous semiconductor laser (809 nm)Room Temp.H2O,
NaOH
1 h99%162[127]
Pd-Au/SiO2AuNRs: ~10 nm (diameter), ~40 nm (length)Suzuki500 W Xe lamp
(>420 nm)
Room Temp.EtOH0.5 h78%334[128]
Au nanorod@Pd superstructuresAuNRs: ~20 nm (diameter), ~60 nm (length)
Pd: 3.8 ± 0.1 nm
Suzuki300 W Xe lamp
(>510 nm)
40 °CEtOH/H2O
(1:3),
NaOH
0.5 h-~2880[129]
Au@Pd NRsAuNRs: 49 ± 5 nm (diameter), 107 ± 8 nm
PdNPs: ~5 nm
SuzukiContinuous 808-nm laser-.H2O,
NaOH
1 h97.6%-[130]
Pd/Au@rGO-10/SiO2AuNRs: 25 nm (diameter), 75 nm (length)
rGO layer: 2.8 nm
Pd: 1.4 nm
Suzuki500 W Xe lamp
(>420 nm)
-EtOH,
K2CO3
0.5 h56%-[131]
GO-Pd@Ag-AgBr>μmSuzuki300 W Xe lamp
(>400 nm)
25 °CEtOH/H2O
(1:1),
K2CO3
0.5 h97%-[132]
Pd/Ag/SBA-15Spherical AgNPs: ~4 nm
Ag nanorods: ~10 nm
Suzuki500 W Xe lamp
(>420 nm)
35 °CEtOH,
K2CO3
6 h53%489[133]
Pd/Ag/SBA-15Pd/Ag NPs: 4.2 nm
SBA-15: >μm
SuzukiXe lampRoom Temp.EtOH
K2CO3
2 h40%-[113]
Supramolecular ensemble-based Ag@Cu2O NPsAgNPs core: 10–15 nm
Cu2O shell: 10–15 nm thickness
Suzuki100 W tungsten filament bulbRoom Temp.EtOH/H2O
(3:1),
K2CO3
5 h75%-[134]
Supramolecular ensemble-based Ag@Cu2O NPsAgNPs core: 7.5 nm
Cu2O shell: 2.5 nm thickness
C-H activation100 W tungsten filament bulbRoom Temp.Toluene/H2O
(3:7),
KOtBu
5.5 h80%-[135]
Ag/TiO2AgNPs: 1.5–5 nm
TiO2: 10-15 nm pores
Suzuki20 W white LED (>420 nm)Room Temp.Toluene24 h97%-[136]
Cu/grapheneCuNPs: ~15 nmC-N cross-coupling300 W Xe lamp
(400–800 nm)
Room Temp.MeOH1 h99%25.4[147]
Supramolecular ensemble 3:CuNPsCuNPs: 3–20 nmC-H alkynylation60 W tungsten filament bulbRoom Temp.DMSO,
K2CO3
8 h72%-[148]
Pd-Cu/SiCPd-Cu NPs: <5 nmSonogashira300 W Xe lamp60 °CDMF,
Cs2CO3
8 h99%-[144]
CuPd@NH2-UiO-66(Zr)CuPd alloy nanoclusters: ~0.9 nmSuzuki300 W Xe lamp
(420–800 nm)
Room Temp.DMF/H2O
(1:1),
TEA
4 h99%-[145]
Cu/Cu2O NPsTetrahexahedron: ~μmUllmannXe lamp--12 h77%-[149]
Pd hexagonal nanoplates60.4 ± 19.3 nm
(20.5 ± 3.7 nm thickness)
SuzukiXe lamp
(300–1000 nm)
25 °CEtOH,
K2CO3
3 h-~288[150]
Cu7S4@PdCu7S4: 14 nm
Pd: 4.3 nm
Suzuki1500-nm diode laserRoom Temp.H2O,
NaOH
0.5 h97%-[151]
Pd/WO3-x NWsPd: ~5 nm
WO3-x NWs: >μm (length), ~10 nm (diameter)
Suzuki500 W Xe lamp
(>650 nm)
-EtOH,
K2CO3
100 min68.75%-[152]
Pd nanoflowers150 nmSuzuki300 W Xe lamp
(>475 nm)
Room Temp.EtOH
Cs2CO3
4 h96%-[153]
Pd/TiO2-Suzuki15 W white LED28 °CH2O-PEG,
NaOC(CH3)3
4 h93%-[154]
Pd/ZnO~25 nmSuzuki11 W white LED lampRoom Temp.H2O,
Cs2CO3
40 min>99%-[155]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shin, H.H.; Suh, Y.D.; Lim, D.-K. Recent Progress in Plasmonic Hybrid Photocatalysis for CO2 Photoreduction and C–C Coupling Reactions. Catalysts 2021, 11, 155. https://doi.org/10.3390/catal11020155

AMA Style

Shin HH, Suh YD, Lim D-K. Recent Progress in Plasmonic Hybrid Photocatalysis for CO2 Photoreduction and C–C Coupling Reactions. Catalysts. 2021; 11(2):155. https://doi.org/10.3390/catal11020155

Chicago/Turabian Style

Shin, Hyeon Ho, Yung Doug Suh, and Dong-Kwon Lim. 2021. "Recent Progress in Plasmonic Hybrid Photocatalysis for CO2 Photoreduction and C–C Coupling Reactions" Catalysts 11, no. 2: 155. https://doi.org/10.3390/catal11020155

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop