Next Article in Journal
Optimization of Naringenin Nanoparticles to Improve the Antitussive Effects on Post-Infectious Cough
Next Article in Special Issue
Competing Metal–Ligand Interactions in Tris(cyclopentadienyl)-cyclohexylisonitrile Complexes of Trivalent Actinides and Lanthanides
Previous Article in Journal
Beyond Moco Biosynthesis―Moonlighting Roles of MoaE and MOCS2
Previous Article in Special Issue
Minimal Active Space for Diradicals Using Multistate Density Functional Theory
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Frustrated Lewis Pairs in Heterogeneous Catalysis: Theoretical Insights

1
State Key Laboratory of Photocatalysis on Energy and Environment, College of Chemistry, Fuzhou University, Fuzhou 350002, China
2
Department of Chemistry and Chemical Biology, University of New Mexico, Albuquerque, NM 87131, USA
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(12), 3734; https://doi.org/10.3390/molecules27123734
Submission received: 9 May 2022 / Revised: 6 June 2022 / Accepted: 7 June 2022 / Published: 10 June 2022
(This article belongs to the Special Issue A Commemorative Special Issue Honoring Professor Donald Truhlar)

Abstract

:
Frustrated Lewis pair (FLP) catalysts have attracted much recent interest because of their exceptional ability to activate small molecules in homogeneous catalysis. In the past ten years, this unique catalysis concept has been extended to heterogeneous catalysis, with much success. Herein, we review the recent theoretical advances in understanding FLP-based heterogeneous catalysis in several applications, including metal oxides, functionalized surfaces, and two-dimensional materials. A better understanding of the details of the catalytic mechanism can help in the experimental design of novel heterogeneous FLP catalysts.

1. Introduction

Noble-metal-based catalysts have been an indispensable part of modern industrial catalysis [1,2,3,4]. However, their high cost and geopolitical risks have always placed a severe constraint on their applications. Considerable efforts have been made to reduce [5,6,7] and/or eliminate the use of noble metals in catalysts [8,9,10]. The former approach has greatly benefited from various strategies and concepts recently developed to improve the atomic efficiency and reactivity of active sites, such as alloyed catalysts [11,12,13], atomic dispersed catalysts [14,15,16], and surface defect engineering [17,18,19]. Meanwhile, significant efforts have been made to achieve similar activity to noble metals by utilizing other materials (e.g., earth-abundant transition metals and main-group elements) [20,21,22,23]. One such strategy is using frustrated Lewis pairs (FLPs), which have attracted much recent attention thanks to their superior performance in homogenous small molecule activation [20,24,25]. As shown in Figure 1, an FLP assembles a Lewis acid (LA) and base (LB) pair in proximity, but with steric hindrance to prevent them from recombining, so as to activate small molecules [26,27,28]. Besides catalyzing heterolytic H2 dissociation [29,30], FLP catalysts have also been used in the activation of other molecules (e.g., alkenes, aldehydes and CO2) [31,32,33,34,35,36,37], thereby providing a new strategy for synthetic chemistry. Inspired by unique FLP-based catalysts in homogenous catalysis, this concept has recently been extended to heterogeneous catalysis [38,39].
Solid surfaces provide a template for designing FLPs, which can be used to aid heterogeneous catalysis [40,41,42]. An early example involved the γ-alumina-catalyzed low-temperature activation of CH4 by Wischert et al. [43], who identified a surface FLP made up of LA and LB sites. The former was the tri-coordinated AlIII, and the latter was O. The dissociation of CH4 led to Al–CH3 and O–H. In another case, Shi et al. observed chain structures on the ZnO(10 1 ¯ 0) surface [44], formed at an extremely low temperature (~20 K) by the heterolytically dissociative adsorption of H2, leading to a zinc hydride and a hydroxyl. Density functional theory (DFT) calculations found two kinds of Zn–O pairs on the surface. One has a bond formed between the Zn and O sites, which can be classified as classical Lewis pairs (CLPs). The other type of Zn–O pair has the Zn and O sites well separated, thereby forming FLPs. DFT calculations suggested that the H2 dissociation on the FLP is more favorable than on the CLP. Similar FLPs have been found on ceria (CeO2) [45,46], which was reported recently to catalyze the selective hydrogenation of unsaturated hydrocarbons [47]. In this case, the FLP, formed between a Ce3+ (LA) created near an oxygen vacancy (Ov) and a nearby oxygen (LB), is shown to catalyze the heterolytic H2 splitting [46]. These surface species have indeed been identified experimentally [48,49], confirming the hypothesis. DFT calculations demonstrated that these species catalyze the subsequent hydrogenation [46].
Driven by these early successes, a large number of theoretical studies have recently been carried out to gain insights into heterogeneous FLP catalysis and to predict new heterogeneous FLP catalysts. The LA and LB sites that make up heterogeneous FLPs may be the electron-deficient and electron-rich atoms and groups externally introduced, or sites with different electronic properties constructed by surface engineering. Among the proposed heterogeneous FLPs, metal-organic frameworks (MOFs) and modified metal surfaces have been studied via density functional theory (DFT) calculations [50,51,52,53], and those based on two-dimensional (2D) materials and metal oxides have started to attract theoretical attention [17,54,55,56]. It is easy to understand why MOFs are good candidates for FLP-based catalysts, as various functional groups can be designed within a MOF with desired distances. On the other hand, metal oxides are naturally rich in LA (metals) and LB sites (oxygens), and surface vacancies can be leveraged in designing FLPs. The designs of FLPs on 2D materials are even more versatile. A typical practice is to use electron-deficient and electron-rich atoms to change the local environments [54,55,57]. These studies have greatly advanced our abilities to activate small molecules with surface FLPs. However, experimental conditions are often more complicated than theoretical models. Although several strategies, mechanisms, and structure–reactivity relationships have been proposed based on these theoretical studies, there is still a lack of consensus on how to design highly efficient FLP catalysts.
Here, we review the recent advances in theoretical studies of FLPs at gas–solid interfaces and discuss their roles in heterogeneous catalysis. We strive to survey the recent theoretical studies of heterogeneous FLP catalysis, aiming to classify various types of heterogeneous FLP catalysts and strategies for creating surface FLPs. We also discuss mechanisms and structure–reactivity relationships that could potentially assist researchers in the rational design of the corresponding catalysts.
This review is organized into four sections. In Section 2, we briefly introduce the original concept of FLPs and the mechanism for small molecule activation by homogeneous FLP catalysts. Several FLP-based heterogeneous catalysts are discussed in Section 3. In Section 3.1, we present our current understanding of several metal-oxide-based FLP catalysts with the relevant catalytic mechanisms. Section 3.2 focuses on the progress of FLP-facilitated catalysis on functionalized surfaces, including MOFs, metal surfaces, and 2D materials. Finally, design strategy, challenges, and outlook are discussed in Section 4.

2. Mechanisms of FLP Catalysis

H2 activation has traditionally been catalyzed by noble metals. In 2006, Stephan and coworkers opened the door to metal-free activation of H2 by introducing an FLP-based homogeneous catalyst (C6H2Me3)2PH(C6F4)BH(C6F5)2 [24], which is shown in Figure 1a. A common form of the FLP consists of an electron-deficient atom (or group) and electron-rich atom (or group) in one molecule or in a molecular pair, which are prevented from neutralizing by utilizing geometry and/or steric hindrance. The role of the FLP is to promote H2 dissociation into protonic (H+) and hydridic (H) species. Since then, homogeneous FLP catalysts have been developed to activate other molecules such as N2 and CO2 [34,58], and much effort has been devoted to the understanding of the mechanism [59,60,61].
Theoretical studies have provided valuable insight into the mechanism of FLP-catalyzed H2 dissociation. The importance of the “frustration” between the LA/LB sites, namely, their spatial/steric separation, was confirmed by a theoretical study of Rokob et al. [62]. The resulting FLP differs from its classical counterpart, CLP, which is stable with low catalytic activity, by creating a pre-organized and strained environment for catalysis, thereby lowering the activation energy for hydrogen splitting, as shown in Figure 2a. Specifically, these authors attributed the catalysis to a covalent transition state involving electron transfer (ET) between the LA/LB pair and the H2 molecule. However, later theoretical studies by Grimme et al. [63] found little evidence for an activated complex LA–H–H–LB and the associated lowered barrier. Instead, the H2 molecule inserted into the FLP cavity dissociates spontaneously, and the activation is essentially the energy cost for H2 to enter the electric field (EF, Figure 2b) generated by the FLP; see Figure 2b. Both the ET and EF mechanisms, which are illustrated in Figure 2c,d, are probably operative to some extent, but their importance might depend on the specific system [64].
The situation in heterogeneous FLP catalysts is expected to be similar to that of its homogeneous counterpart. The possible mechanisms and corresponding applications of FLPs in heterogeneous catalysis are discussed in detail in the following sections.

3. Current Developments in Heterogeneous FLP Catalysts

3.1. FLPs Based on Metal Oxides

Metal oxides, possessing natural LA and LB sites on their surfaces, are known as heterogeneous catalysts for C–H and H–H bond activation. Studies suggested that the Lewis acid/base properties of the active sites on metal oxide surfaces could provide a unique way to obtain an in-depth understanding of the catalysis [65].
Λ-alumina (γ-Al2O3) is considered as a heterogeneous FLP catalyst for C–H bond activation in CH4. In a combined experimental and first-principles study, Wischert et al. revealed that low-temperature heterolytic C–H bond cleavage in CH4 on a hydroxylated γ-Al2O3(110) to form Al-CH3 and O-H can be ascribed to the co-action of non-adjacent Lewis acid–base pairs consisting of surface AlIII (LA) and O (LB) sites [43]. These two sites are separated by 4.1 Å on the hydroxylated alumina surface, forming a cavity for the adsorption and activation of CH4, leading to Al–CH3 and O–H. They also speculated that such FLP catalysis might be general for different oxides towards other molecules with polarizable X+–Y bonds. In the subsequent study by the same group, the authors theoretically compared the C−H or H−H bond cleavage of CH4 or H2 catalyzed by a frustrated Al–O pair over the γ-Al2O3(110) and γ-Al2O3(100) surfaces as a function of hydroxyl coverage [66]; see Figure 3. Similarly to the case of CH4 activation, the activation of the H–H bond in H2 leads to Al−H and O–H species. The activation of H2 is more facile because of the lower dissociation barrier than that for CH4 activation. Importantly, these studies demonstrated that the reactivity of Al and O sites is affected by partial hydroxylation, which stabilizes the metastable (110) surfaces and tunes the electronic structures of the Al–O FLP by increasing the acidity of Al and the basicity of O. Hence, surface hydration can be used as an effective strategy to create and stabilize surface FLPs on (nonreducible) metal oxide surfaces.
Another example of FLP-facilitated heterolytic dissociation of H2 was reported to occur on hydroxylated indium oxide (In2O3−x(OH)y) which catalyzes the reverse water gas shift (RWGS) reaction (CO2 + H2  CO + H2O). Starting with the H2 activation on the surface, two dissociated H atoms and one of O atoms from CO2 molecules combine to form a H2O molecule, leaving CO. In DFT calculations of this reaction combined with in situ spectroscopic and kinetic studies by Ghuman et al. [67], the activation of H2 was attributed to surface FLPs consisting of non-adjacent hydroxide (LB) and indium (LA) pairs, and the later are exposed at a surface oxygen vacancy. The proposed mechanism leads to the formation of surface O–H2 and In–H species, which subsequently attack the adsorbed CO2 to produce CO and H2O.
The existence of FLPs is of course not restricted to the two examples discussed above. Ceria (CeO2) has long been used as a support for metal catalysts, thanks to its strong reducibility [68]. The recent discovery by Vilé et al. that ceria alone is capable of catalyzing selective hydrogenation of unsaturated hydrocarbons [47,69,70] has stimulated considerable interest in understanding its reactivity and selectivity [71]. The initial proposal of the mechanism for partial acetylene hydrogenation contained a relatively high barrier (2.86 eV) [69], which is kinetically unattainable. The key flaw of that mechanism was the assumption of homolytic activation of H2, leading to the formation of two surface OH species, which are responsible for the high barrier in the hydrogenation step. Interestingly, other DFT studies showed that heterolytic activation of H2 has a significantly lower barrier [72]. In this case, an exposed Ce serves as the LA for accepting the hydride (H−) species from the dissociating H2, and a surface O can serve as the LB for the proton (H+). Subsequent DFT calculations revealed the crucial role of O vacancies in forming the surface FLPs [45,73], and mechanism investigation suggests that the energy barrier of the rater-determining step of acetylene hydrogenation catalyzed by the FLP is lower than that of a CLP on CeO2(110) facets (Figure 4a). One can see from the charge density difference (CDD) in Figure 4b that in the TS of H2 activation, electron density is increased around H on Ce and decreased around H on O with both CLP and FLP. Meanwhile, it can be inferred from the electron localization function (ELF) maps that there are some highly delocalized regions between H- and Ce, and the region between O and H+ shows a strong covalent property. All these suggested that both the local electric field on the surface and the electron transfer between H2 and FLP/CLP contribute the reactivity. Based on our DFT calculations on CeO2(111), the proton and hydride species are crucial for the subsequent hydrogenation steps [46,74], leading to a mechanism that is consistent with experimental observations [49]. All the above studies suggested that the surface FLPs are crucial for the heterolytic H2 dissociation. Recently, Wu et al., using in situ inelastic neutron scattering spectroscopy (INS), revealed that the hydride (Ce–H) species is indeed present during the acetylene hydrogenation catalyzed by ceria [48]. In the meantime, Moon et al. detected surface OH species on the ceria catalyst using in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) [75]. These experimental data thus provided strong supporting evidence for the heterolytic dissociation mechanism catalyzed by surface FLPs.
Doping has been shown to enhance the FLP’ catalytic activity on ceria. For example, Ni doping promotes oxygen vacancies (Figure 4c), leading to effective Ce/O FLPs [46,76] without directly participating in catalysis. Doping of Ga can also lead to novel single-atom catalysts in which the LA is provided by the doped metal on ceria surfaces [74], and results in a lower barrier for H2 activation and more active H species, as illustrated by Figure 4d. Other dopants include copper [77,78]. DFT results show that atomic Cu promotes the formation of Ov [71], leading to Cu/O FLPs catalyzing the cleavage of the H–H bond. These Ce–O FLPs were also found to be active for CH4 activation, thanks to the strong interaction between the CH4 molecule and FLPs. In addition, these FLPs were shown to efficiently catalyze the non-oxidative coupling of CH4 to C2H6 and C2H4 [79].
We noticed a recent DFT study by Lin’s group showing that Ti and O sites on anatase TiO2(100) and TiO2(110) surfaces can form diverse FLPs with different distances between Ti and O (Figure 5a) in the presence of oxygen vacancies [80]. These Ti–O FLPs can heterolytically cleave H2 to form Ti–H and O–H species, which then react with acetylene to produce ethylene. Interestingly, the calculation results showed a volcano-shaped relationship between the H2 activation energy and the strength of the O–H bond (Figure 5b), and the barriers of the first and second steps of acetylene hydrogenation are positively correlated with the strength of Ti–H bonds and C2H3 adsorption energy, respectively. Importantly, this theoretical prediction of the hydrogenation reactivity of TiO2 was confirmed by an experiment [80], thereby providing useful insights for the future development of metal-oxide-based FLPs catalysts.
It is important to recognize that the FLP is a dynamic entity whose properties can be altered by external parameters such as temperature. Indeed, the impact of temperature on the surfaces of In/O FLPs with regard to catalyzing RWGS was investigated in the temperature range of 20 to 180 °C by using the metadynamics-biased ab initio molecular dynamics (metaD AIMD) [81]. The FLPs were found to be structurally altered at a high temperature (180 °C), as the distance between the LA and LB fluctuated by 0.04 Å. This led to a reduction of the barrier of the heterolytic H2 dissociation by 0.15 eV (based on Perdew−Burke−Ernzerhof (PBE) with Rappe−Rabe−Kaxiras−Joannopoulos (RRKJ), via Quantum Espresso (QE)) compared to that at 20 °C (Figure 6a). While the reduction of CO2, considered as the rate-limiting step, is not sensitive to temperature changes, the calculated energy barrier for the adsorption of CO2 at the FLPs is reduced by 0.19 eV (Figure 6b). The results highlight the important role of thermal fluctuation in the spatial separation of FLPs, which is closely related to the enhanced reactivity.
In another example, Fabris and workers computed the free energy surface for H2 activation on CeO2(111) using metadynamics [82]. Although no oxygen vacancy was included in the simulation, this study also demonstrated that the barrier (0.73 eV, based on PBE via QE combined PLUMED) is significantly reduced at the simulation temperature (350 K) relative to the 0 K result (0.99 eV), due apparently to the thermal fluctuation of the surface O species.
The temperature dependence of stable Ce–O FLPs on CeO2(110) was discussed by Huang et al., who reported that the formation of surface FLPs is dependent on the number of oxygen vacancies [79]. Although static DFT calculations indicated that the FLPs in the presence of fewer oxygen vacancies were unstable, AIMD results showed that the FLP sites can dynamically regenerate at high temperatures.

3.2. FLPs Based on Functionalized Surfaces

As discussed in the above section, heterogeneous FLPs on metal oxides are assembled from pre-existing Lewis acid and base sites. However, for most solid surfaces, the chemical environment is relatively simple and does not allow the natural formation of FLPs. Therefore, the introduction of foreign species as the LA or LB on the surface is an effective strategy for constructing surface FLPs. These studies not only provide us with alternative perspectives to understand the processes involved in heterogeneous catalysis, but also extend the application of heterogeneous FLP catalysis and design strategies for future FLP-based catalysts.
The simplest case would be pre-absorption of LAs and LBs on surfaces as the electron acceptor or donor, respectively. Lu et al. found that the pre-adsorption of imine or nitrile as LBs can enable efficient H2 activation on a modified gold surface, where the gold atom serves as the LA [50] (Figure 7a). The DFT calculation results suggested that the enhanced reactivity of the functionalized gold surface can be ascribed to the synergetic effects of the frustrated LA and LB sites, on which the partially filled s- and p-states of gold accept the electrons from the H2 σ orbital (Figure 7b). Similarly, Fiorio et al. reported that the pre-adsorption of nitrogen-containing Lewis bases on Au nanoparticles can also couple with surface Au atoms to generate heterogeneous FLPs [51,52]. Interestingly, they found that FLPs composed of Au LA sites and amine species with different basicity can be used to tune H2 heterolytic activation, which is consistent with the experimental observations. In addition, a reaction intermediate can also act as LB. For example, in a first-principles study, Jian et al. demonstrated that Ni1(OH)2/TiO2 possesses superior reactivity and selectivity for acetylene hydrogenation [83], in which the adsorbed C2H3 and C2H5 intermediates produced during the acetylene hydrogenation can serve as LBs to form FLPs with the surface Ni sites to accelerate the heterolytic H2 cleavage. The enhanced reactivity of metal surfaces by the introduction of foreign species to create new FLPs offers a useful strategy for the development of other metal-based hydrogenation catalysts.
Porous MOFs provide a flexible template and suitable nanospace to create FLPs. For example, Ye et al. reported that 1-(difluoroboranyl)-4-methyl-1H-pyrazole grafted in the MOF of UiO-66 exhibits the characteristics of an FLP formed by P (LB) and B (LA) atoms [84]. From first-principles calculations, the authors found that this FLP has the capacity to catalyze the heterolytic H2 dissociation to form P–H and B–H species, the key step in CO2 reduction with H2. The calculated minimal energy reaction pathway indicates that the CO2 is hydrogenated with the pre-adsorbed H species at FLP sites. The heterogeneous FLP (UiO-66-P-BF2) showed higher reactivity than the homogeneous FLPs catalyst (1-[bis(pentafluorophenyl)boryl]-3,5-aditert-butyl-1H-pyrazole).
In addition, FLPs can also be created via engineering the interface between metal and metal oxide. For example, Zhao et al. reported that interfaces between metallic or oxidized nickel species and a Au surface (Figure 8a) are active for H2O dissociation [85]. Furthermore, their DFT results showed that the Ni13O13/Au(111), Ni10/Au(111), Ni10O6/Au(111), and Ni10O6-Ni6/Au(111) interfaces have different adsorption strengths for the adsorption of H2O (Figure 8a). Among them, the NiOx–Ni interface possessed the strongest adsorption of H2O, which is controlled by the Lewis acidity of the Ni site. In particular, H2O dissociation at the NiOx-Ni interface is a spontaneous process (Figure 8b) in which metallic Ni acts as an LA to accept the hydroxyl, while the O serves as the LB to capture a proton with the frustrated LA–LB distance of 3.41 Å (Figure 8c). This work points out that the heterogeneous FLPs can also be constructed at metal and metal oxide interfaces, extending the applications of FLP in heterogeneous catalysis.
Besides the surface FLPs mentioned above, FLPs based on nanostructured carbon materials have also been reported recently. Primo et al. demonstrated that graphene can hydrogenate the C2H2 to C2H4 with high conversion rates and selectivity in the absence of noble metals [86]. Interestingly, the addition of metal impurities (e.g., Mn and Pd) affects the hydrogenation performance little, and even sometimes decreases the reactivity. The comparison between the performance of graphene, graphene oxide, reduced graphene oxide, and N/P/S-contained graphene also suggested the reactivity originates from intrinsic sites of graphene. More interestingly, the reactivity of graphene was significantly and reversibly affected by the pre-adsorption of CO2 and NH3. Based on these observations, Primo et al. concluded that the hydrogenation reactivity can be ascribed to surface FLPs. However, the mechanism from the perspective of FLPs is still unexplained.
Of course, FLPs can also be created by substitutive doping, which changes the chemical structure of graphene. Sun et al. proposed an FLP catalyst for H2 activation by boron and nitrogen co-doping the bilayer graphene (BN-G) [87]. From DFT calculations, three BN-G configurations were obtained, including AA stacking (AA-01), AB stacking (AB-01), and AB stacking (AB-02). H2 activation in the interface between the two layers of these three structures was further investigated and compared with the process catalyzed by pristine bilayer graphene. The calculated activation energy of H2 dissociation on pristine bilayer graphene is 2.3 eV; it is 0.99, 1.05, and 1.36 eV on AA-01, AB-01, and AB-02, respectively (TPSS function with D2 via VASP). It was thus proposed that the FLP formed by B and N catalyzes the heterolytic H2 cleavage. Furthermore, Sun et al. found that dopants at the graphene edge possess higher reactivity towards H2 dissociation, and that H2 could spontaneously dissociate at the B site on the edge, revealing that the dopants at the edge could contribute more reactivity than the dopants in the basal plane. Very recently, Chen et al. argued that these B–N sites might not be responsible for these FLPs [56].
Besides 2D carbon-based materials, other 2D structures have also been investigated for constructing FLP-based catalysts. Zhao et al. reported an FLP catalyst based on phosphorene via a doping strategy [54]. In particular, the B introduced to the phosphorene monolayer serves as the LA and P acts as the LB (Figure 9a, left). The charge transfer can be seen in the ELF results (Figure 9a, right): the electron density on the B atom is highly delocalized, indicating it is charge-deficient, whereas the electron density on the P atom is strongly localized, suggesting that it has lone pair electrons. H2 activation catalyzed by such an FLP would result in a much lower barrier of 0.59 eV (based on double numerical plus polarization basis set and PBE+D2, via DMol3) (Figure 9b) compared to 2.25 eV on pristine phosphorene, suggesting high activity of the newly formed FLPs for H2 activation. They also found that, using Me2C=O, HCN, and C2H4 as probe molecules, the H species produced over the FLPs can attack the unsaturated molecules to complete the hydrogenation catalysis.
Doping of 2D materials by metals can also lead to FLPs. Chen et al. reported transition metals and boron co-doped phosphorene with promising N2 reduction reaction (NRR) electrochemical performance [88]. Usually, it is well recognized that the electrocatalytic NRR performance can be improved by stabilizing the key N2H intermediates [89]. These authors demonstrated an alternative strategy based on the concept of FLPs. Accordingly, the B atom acts as the LA to provide an empty orbital, while the transition metal acts as the LB to donate electrons for enhancing N2H adsorption.
The carbon nitride group is another 2D material that is widely used in heterogeneous catalysis [90]. For example, Wan et al. reported Al/B doping C2N and g-C3N4 to construct FLPs from their DFT calculations [55]. These theoretical results showed that Al–N FLPs were more active than B–N FLPs for H2 activation. Introducing Al as the LA breaks the electron-rich environments at the pore edge and forms several Al–N combinations with varying electronic structures as FLPs (Figure 10a). These FLPs could efficiently activate H2 to form Al–H and N–H (Figure 10b), which can then hydrogenate C2H2 to C2H4. Similarly to transition states on defect CeO2(110) (Figure 4b), CDD and ELF maps in Figure 10c suggest the electron transfer between H2 with the Al/N FLP. Both cases utilize the intrinsic LB sites of carbon nitride, taking advantage of the introduced electron-deficient LA, to achieve the small molecule activation in the absence of noble metals.

4. Conclusions and Prospects

Significant efforts have recently been made to explore the applications of heterogeneous FLP catalysts for various reactions. Herein, we provided a comprehensive review of theoretical understanding of the mechanisms of FLP-facilitated heterogeneous catalysis. The insights provided by these theoretical results, coupled with experimental observations, offer a framework for in-depth understandings of heterogeneous FLP catalysis and design principles.
The key to heterogeneous FLP catalysis is the presence of spatially separated but adjacent surface active sites with electron-rich and deficient properties, respectively. The geometric separation of these LA and LB sites is important because it prevents direct bonding that would neutralize them, as in a CLP, while offering a suitable cavity for the cleavage of the targeted bond. Hence, the optimization of the separation between the FLP sites and the acidity/basicity is essential for the catalyzed activation of molecules, such as H2 and CH4. The adjacent LA and LB sites also polarize the molecule, resulting in a lowered dissociation barrier. In the H2 case, for example, its dissociation results in the formation of protonic and hydridic species, which can then be used to carry out subsequent reaction steps.
The heterogeneous FLP systems reported so far can be generally divided into two categories based on materials. One class is based on metal oxides, and the other relies on functionalizing surfaces. The former has abundant LA(M) and LB(O) sites of its own and is subjected to modifications by external means, such as substitutive doping. The doped metal can either serve as the LA or help to create an LA site by promoting oxygen vacancies. Alternatively, one can also control the LB site by hydroxylation. On the other hand, the latter type relies on the introduction of LA and LB species through functionalization. This can be readily achieved in porous materials such as MOFs, where the distance between the LA and LB sites can be readily controlled. The functionalization can also take place on metal surfaces, where the introduction of LB species transforms metal atoms as LA sites. Non-metal FLPs can be created analogously by doping 2D materials such as graphene. These strategies provide a range of tools for designing FLPs in heterogeneous systems.
The rapid development in the field serves as a manifestation of the powerfulness and universality of the FLP concept in heterogeneous catalysis. We expect more flexible ways to create FLPs in the future with traditional and novel LAs and LBs. Materials, such as metal-support interfaces and corrugated/reconstructed 2D materials, possessing regions with different electron gaining and losing abilities, might be leveraged to generate potential FLPs. Due to its outstanding performance in activating small molecules and the avoidance of noble metals, FLPs have prospects in wider applications in the future for heterogeneous catalysis. We expect robust future development in this field.

Author Contributions

Conceptualization, S.L. and H.G.; writing—original draft preparation, Q.W.; writing—review and editing, S.L. and H.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (21973013 to S.L.), the Natural Science Foundation of Fujian Province of China (2020J02025 to S.L.), and the “Chuying Program” for the Top Young Talents of Fujian Province to S.L. H.G. acknowledges support from the National Science Foundation (CHE-1951328) and a Humboldt Research Award from the Alexander von Humboldt Foundation.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brown, M.; Green, A.; Cohn, G.; Andersen, H.J.I.; Chemistry, E. Purifying hydrogen by selective oxidation of carbon monoxide. Ind. Eng. Chem. 1960, 52, 841–844. [Google Scholar] [CrossRef]
  2. Oh, S.H.; Sinkevitch, R.M. Carbon Monoxide Removal from Hydrogen-Rich Fuel Cell Feedstreams by Selective Catalytic Oxidation. J. Catal. 1993, 142, 254–262. [Google Scholar] [CrossRef]
  3. Rosso, I.; Galletti, C.; Saracco, G.; Garrone, E.; Specchia, V. Development of A zeolites-supported noble-metal catalysts for CO preferential oxidation: H2 gas purification for fuel cell. Appl. Catal. B 2004, 48, 195–203. [Google Scholar] [CrossRef]
  4. Miyata, T.; Li, D.; Shiraga, M.; Shishido, T.; Oumi, Y.; Sano, T.; Takehira, K. Promoting effect of Rh, Pd and Pt noble metals to the Ni/Mg(Al)O catalysts for the DSS-like operation in CH4 steam reforming. Appl. Catal. A 2006, 310, 97–104. [Google Scholar] [CrossRef]
  5. Ren, N.; Yang, Y.-H.; Shen, J.; Zhang, Y.-H.; Xu, H.-L.; Gao, Z.; Tang, Y. Novel, efficient hollow zeolitically microcapsulized noble metal catalysts. J. Catal. 2007, 251, 182–188. [Google Scholar] [CrossRef]
  6. Jin, Z.; Zhang, X.; Lu, G.; Li, S. Improved quantum yield for photocatalytic hydrogen generation under visible light irradiation over eosin sensitized TiO2—Investigation of different noble metal loading. J. Mol. Catal. A Chem. 2006, 259, 275–280. [Google Scholar] [CrossRef]
  7. Bai, S.; Liu, F.; Huang, B.; Li, F.; Lin, H.; Wu, T.; Sun, M.; Wu, J.; Shao, Q.; Xu, Y.; et al. High-efficiency direct methane conversion to oxygenates on a cerium dioxide nanowires supported rhodium single-atom catalyst. Nat. Commun. 2020, 11, 954. [Google Scholar] [CrossRef]
  8. Yu, D.; Nagelli, E.; Du, F.; Dai, L. Metal-Free Carbon Nanomaterials Become More Active than Metal Catalysts and Last Longer. J. Phys. Chem. Lett. 2010, 1, 2165–2173. [Google Scholar] [CrossRef]
  9. Lu, S.; Pan, J.; Huang, A.; Zhuang, L.; Lu, J. Alkaline polymer electrolyte fuel cells completely free from noble metal catalysts. Proc. Natl. Acad. Sci. USA 2008, 105, 20611. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76–80. [Google Scholar] [CrossRef]
  11. Wang, A.-Q.; Liu, J.-H.; Lin, S.D.; Lin, T.-S.; Mou, C.-Y. A novel efficient Au–Ag alloy catalyst system: Preparation, activity, and characterization. J. Catal. 2005, 233, 186–197. [Google Scholar] [CrossRef]
  12. Xu, J.; White, T.; Li, P.; He, C.; Yu, J.; Yuan, W.; Han, Y.-F. Biphasic Pd–Au Alloy Catalyst for Low-Temperature CO Oxidation. J. Am. Chem. Soc. 2010, 132, 10398–10406. [Google Scholar] [CrossRef] [PubMed]
  13. Greeley, J.; Mavrikakis, M. Alloy catalysts designed from first principles. Nat. Mater. 2004, 3, 810–815. [Google Scholar] [CrossRef] [PubMed]
  14. Qiao, B.; Wang, A.; Yang, X.; Allard, L.F.; Jiang, Z.; Cui, Y.; Liu, J.; Li, J.; Zhang, T. Single-atom catalysis of CO oxidation using Pt1/FeOx. Nat. Chem. 2011, 3, 634–641. [Google Scholar] [CrossRef]
  15. Wang, L.; Zhang, W.; Wang, S.; Gao, Z.; Luo, Z.; Wang, X.; Zeng, R.; Li, A.; Li, H.; Wang, M.; et al. Atomic-level insights in optimizing reaction paths for hydroformylation reaction over Rh/CoO single-atom catalyst. Nat. Commun. 2016, 7, 14036. [Google Scholar] [CrossRef] [Green Version]
  16. Feng, J.; Gao, H.; Zheng, L.; Chen, Z.; Zeng, S.; Jiang, C.; Dong, H.; Liu, L.; Zhang, S.; Zhang, X. A Mn-N3 single-atom catalyst embedded in graphitic carbon nitride for efficient CO2 electroreduction. Nat. Commun. 2020, 11, 4341. [Google Scholar] [CrossRef]
  17. Wang, Q.; Lei, Y.; Wang, D.; Li, Y. Defect engineering in earth-abundant electrocatalysts for CO2 and N2 reduction. Energy Environ. Sci. 2019, 12, 1730–1750. [Google Scholar] [CrossRef]
  18. Ortiz-Medina, J.; Wang, Z.; Cruz-Silva, R.; Morelos-Gomez, A.; Wang, F.; Yao, X.; Terrones, M.; Endo, M. Defect Engineering and Surface Functionalization of Nanocarbons for Metal-Free Catalysis. Adv. Mater. 2019, 31, 1805717. [Google Scholar] [CrossRef]
  19. Yue, Q.; Liu, C.; Wan, Y.; Wu, X.; Zhang, X.; Du, P. Defect engineering of mesoporous nickel ferrite and its application for highly enhanced water oxidation catalysis. J. Catal. 2018, 358, 1–7. [Google Scholar] [CrossRef]
  20. Geier, S.J.; Stephan, D.W. Lutidine/B(C6F5)3: At the Boundary of Classical and Frustrated Lewis Pair Reactivity. J. Am. Chem. Soc. 2009, 131, 3476–3477. [Google Scholar] [CrossRef]
  21. Simanullang, W.F.; Itahara, H.; Takahashi, N.; Kosaka, S.; Shimizu, K.-i.; Furukawa, S. Highly active and noble-metal-alternative hydrogenation catalysts prepared by dealloying Ni–Si intermetallic compounds. Chem. Commun. 2019, 55, 13999–14002. [Google Scholar] [CrossRef]
  22. Liang, Q.; Hayashi, K.; Song, D. Catalytic Alkyne Dimerization without Noble Metals. ACS Catal. 2020, 10, 4895–4905. [Google Scholar] [CrossRef]
  23. Basu, M.; Zhang, Z.-W.; Chen, C.-J.; Chen, P.-T.; Yang, K.-C.; Ma, C.-G.; Lin, C.C.; Hu, S.-F.; Liu, R.-S. Heterostructure of Si and CoSe2: A Promising Photocathode Based on a Non-noble Metal Catalyst for Photoelectrochemical Hydrogen Evolution. Angew. Chem. Int. Ed. 2015, 54, 6211–6216. [Google Scholar] [CrossRef] [PubMed]
  24. Welch, G.C.; Juan, R.R.S.; Masuda, J.D.; Stephan, D.W. Reversible, Metal-Free Hydrogen Activation. Science 2006, 314, 1124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Chase, P.A.; Stephan, D.W. Hydrogen and Amine Activation by a Frustrated Lewis Pair of a Bulky N-Heterocyclic Carbene and B(C6F5)3. Angew. Chem. Int. Ed. 2008, 47, 7433–7437. [Google Scholar] [CrossRef] [PubMed]
  26. Stephan, D.W.; Erker, G. Frustrated Lewis pairs: Metal-free hydrogen activation and more. Angew. Chem. Int. Ed. 2010, 49, 46–76. [Google Scholar] [CrossRef] [PubMed]
  27. Stephan, D.W. The broadening reach of frustrated Lewis pair chemistry. Science 2016, 354, aaf7229. [Google Scholar] [CrossRef]
  28. Jupp, A.R.; Stephan, D.W. New Directions for Frustrated Lewis Pair Chemistry. Trends Chem. 2019, 1, 35–48. [Google Scholar] [CrossRef]
  29. Welch, G.C.; Stephan, D.W. Facile Heterolytic Cleavage of Dihydrogen by Phosphines and Boranes. J. Am. Chem. Soc. 2007, 129, 1880–1881. [Google Scholar] [CrossRef]
  30. Spikes, G.H.; Fettinger, J.C.; Power, P.P. Facile Activation of Dihydrogen by an Unsaturated Heavier Main Group Compound. J. Am. Chem. Soc. 2005, 127, 12232–12233. [Google Scholar] [CrossRef]
  31. McCahill, J.S.J.; Welch, G.C.; Stephan, D.W. Reactivity of “Frustrated Lewis Pairs”: Three-Component Reactions of Phosphines, a Borane, and Olefins. Angew. Chem. Int. Ed. 2007, 46, 4968–4971. [Google Scholar] [CrossRef] [PubMed]
  32. Ullrich, M.; Seto, K.S.H.; Lough, A.J.; Stephan, D.W. 1,4-Addition reactions of frustrated Lewis pairs to 1,3-dienes. Chem. Commun. 2009, 17, 2335–2337. [Google Scholar] [CrossRef] [PubMed]
  33. Mömming, C.M.; Otten, E.; Kehr, G.; Fröhlich, R.; Grimme, S.; Stephan, D.W.; Erker, G. Reversible Metal-Free Carbon Dioxide Binding by Frustrated Lewis Pairs. Angew. Chem. Int. Ed. 2009, 48, 6643–6646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Jiang, Y.; Blacque, O.; Fox, T.; Berke, H. Catalytic CO2 Activation Assisted by Rhenium Hydride/B(C6F5)3 Frustrated Lewis Pairs—Metal Hydrides Functioning as FLP Bases. J. Am. Chem. Soc. 2013, 135, 7751–7760. [Google Scholar] [CrossRef]
  35. Pal, R.; Ghara, M.; Chattaraj, P.K. Activation of Small Molecules and Hydrogenation of CO2 Catalyzed by Frustrated Lewis Pairs. Catalysts 2022, 12, 201. [Google Scholar] [CrossRef]
  36. Ghara, M.; Giri, S.; Das, P.; Chattaraj, P. Possible C-F Bond Activation by B(C6F5)3/Lutidine and Al(C6F5)3/Lutidine Frustrated Lewis Pair: An In Silico Study. J. Chem. Sci. 2021, 134, 14. [Google Scholar] [CrossRef]
  37. Sheng, J.; He, Y.; Huang, M.; Yuan, C.; Wang, S.; Dong, F. Frustrated Lewis Pair Sites Boosting CO2 Photoreduction on Cs2CuBr4 Perovskite Quantum Dots. ACS Catalysis 2022, 12, 2915–2926. [Google Scholar] [CrossRef]
  38. Liu, S.; Dong, M.; Wu, Y.; Luan, S.; Xin, Y.; Du, J.; Li, S.; Liu, H.; Han, B. Solid surface frustrated Lewis pair constructed on layered AlOOH for hydrogenation reaction. Nat. Commun. 2022, 13, 2320. [Google Scholar] [CrossRef]
  39. Xing, J.-Y.; Buffet, J.-C.; Rees, N.H.; Nørby, P.; O’Hare, D. Hydrogen cleavage by solid-phase frustrated Lewis pairs. Chem. Commun. 2016, 52, 10478–10481. [Google Scholar] [CrossRef]
  40. Ma, Y.; Zhang, S.; Chang, C.-R.; Huang, Z.-Q.; Ho, J.C.; Qu, Y. Semi-solid and solid frustrated Lewis pair catalysts. Chem. Soc. Rev. 2018, 47, 5541–5553. [Google Scholar] [CrossRef]
  41. Jupp, A.R. Heterogeneous Catalysis by Frustrated Lewis Pairs. In Frustrated Lewis Pairs; Chris Slootweg, J., Jupp, A.R., Eds.; Springer International Publishing: Cham, Switzerland, 2021; pp. 237–281. [Google Scholar]
  42. Zhang, Y.; Lan, P.C.; Martin, K.; Ma, S. Porous frustrated Lewis pair catalysts: Advances and perspective. Chem. Catal. 2022, 2, 439–457. [Google Scholar] [CrossRef]
  43. Wischert, R.; Coperet, C.; Delbecq, F.; Sautet, P. Optimal water coverage on alumina: A key to generate Lewis acid-base pairs that are reactive towards the C-H bond activation of methane. Angew. Chem. Int. Ed. 2011, 50, 3202–3205. [Google Scholar] [CrossRef] [PubMed]
  44. Shi, H.; Yuan, H.; Li, Z.; Wang, W.; Li, Z.; Shao, X. Low-Temperature Heterolytic Adsorption of H2 on ZnO(10 10 ¯ ) Surface. J. Phys. Chem. C 2019, 123, 13283–13287. [Google Scholar] [CrossRef]
  45. Huang, Z.-Q.; Liu, L.-P.; Qi, S.; Zhang, S.; Qu, Y.; Chang, C.-R. Understanding all-solid frustrated-Lewis-pair sites on CeO2 from theoretical perspectives. ACS Catal. 2018, 8, 546–554. [Google Scholar] [CrossRef]
  46. Riley, C.; Zhou, S.; Kunwar, D.; De La Riva, A.; Peterson, E.; Payne, R.; Gao, L.; Lin, S.; Guo, H.; Datye, A. Design of Effective Catalysts for Selective Alkyne Hydrogenation by Doping of Ceria with a Single-Atom Promotor. J. Am. Chem. Soc. 2018, 140, 12964–12973. [Google Scholar] [CrossRef] [PubMed]
  47. Vilé, G.; Bridier, B.; Wichert, J.; Pérez-Ramírez, J. Ceria in Hydrogenation Catalysis: High Selectivity in the Conversion of Alkynes to Olefins. Angew. Chem. Int. Ed. 2012, 51, 8620–8623. [Google Scholar] [CrossRef] [PubMed]
  48. Wu, Z.; Cheng, Y.; Tao, F.; Daemen, L.; Foo, G.S.; Nguyen, L.; Zhang, X.; Beste, A.; Ramirez-Cuesta, A.J. Direct neutron spectroscopy observation of cerium hydride species on a cerium oxide catalyst. J. Am. Chem. Soc. 2017, 139, 9721–9727. [Google Scholar] [CrossRef]
  49. Werner, K.; Weng, X.; Calaza, F.; Sterrer, M.; Kropp, T.; Paier, J.; Sauer, J.; Wilde, M.; Fukutani, K.; Shaikhutdinov, S.; et al. Toward an understanding of selective alkyne hydrogenation on ceria: On the impact of O vacancies on H2 interaction with CeO2(111). J. Am. Chem. Soc. 2017, 139, 17608–17616. [Google Scholar] [CrossRef]
  50. Lu, G.; Zhang, P.; Sun, D.; Wang, L.; Zhou, K.; Wang, Z.-X.; Guo, G.-C. Gold catalyzed hydrogenations of small imines and nitriles: Enhanced reactivity of Au surface toward H2via collaboration with a Lewis base. Chem. Sci. 2014, 5, 1082–1090. [Google Scholar] [CrossRef]
  51. Fiorio, J.L.; López, N.; Rossi, L.M. Gold–Ligand-Catalyzed Selective Hydrogenation of Alkynes into cis-Alkenes via H2 Heterolytic Activation by Frustrated Lewis Pairs. ACS Catal. 2017, 7, 2973–2980. [Google Scholar] [CrossRef] [Green Version]
  52. Fiorio, J.L.; Gonçalves, R.V.; Teixeira-Neto, E.; Ortuño, M.A.; López, N.; Rossi, L.M. Accessing Frustrated Lewis Pair Chemistry through Robust Gold@N-Doped Carbon for Selective Hydrogenation of Alkynes. ACS Catal. 2018, 8, 3516–3524. [Google Scholar] [CrossRef]
  53. Shyshkanov, S.; Nguyen, T.N.; Ebrahim, F.M.; Stylianou, K.C.; Dyson, P.J. In Situ Formation of Frustrated Lewis Pairs in a Water-Tolerant Metal-Organic Framework for the Transformation of CO2. Angew. Chem. Int. Ed. 2019, 58, 5371–5375. [Google Scholar] [CrossRef] [PubMed]
  54. Zhao, J.; Liu, X.; Chen, Z. Frustrated Lewis Pair Catalysts in Two Dimensions: B/Al-Doped Phosphorenes as Promising Catalysts for Hydrogenation of Small Unsaturated Molecules. ACS Catal. 2016, 7, 766–771. [Google Scholar] [CrossRef]
  55. Wan, Q.; Li, J.; Jiang, R.; Lin, S. Construction of frustrated Lewis pairs on carbon nitride nanosheets for catalytic hydrogenation of acetylene. Phys. Chem. Chem. Phys. 2021, 23, 24349–24356. [Google Scholar] [CrossRef] [PubMed]
  56. Chen, H.; Xiong, C.; Moon, J.; Ivanov, A.S.; Lin, W.; Wang, T.; Fu, J.; Jiang, D.-e.; Wu, Z.; Yang, Z.; et al. Defect-Regulated Frustrated-Lewis-Pair Behavior of Boron Nitride in Ambient Pressure Hydrogen Activation. J. Am. Chem. Soc. 2022, in press. [CrossRef]
  57. Ran, Y.; Yu, X.; Liu, J.; Cui, J.; Wang, J.; Wang, L.; Zhang, Y.; Xiang, X.; Ye, J. Polymeric carbon nitride with frustrated Lewis pair sites for enhanced photofixation of nitrogen. J. Mater. Chem. A 2020, 8, 13292–13298. [Google Scholar] [CrossRef]
  58. Rouf, A.M.; Dai, C.; Xu, F.; Zhu, J. Dinitrogen Activation by Tricoordinated Boron Species: A Systematic Design. Adv. Theory Simul. 2020, 3, 1900205. [Google Scholar] [CrossRef]
  59. Daru, J.; Bakó, I.; Stirling, A.; Pápai, I. Mechanism of Heterolytic Hydrogen Splitting by Frustrated Lewis Pairs: Comparison of Static and Dynamic Models. ACS Catal. 2019, 9, 6049–6057. [Google Scholar] [CrossRef] [Green Version]
  60. Liu, L.; Lukose, B.; Jaque, P.; Ensing, B. Reaction mechanism of hydrogen activation by frustrated Lewis pairs. Green Energy Environ. 2019, 4, 20–28. [Google Scholar] [CrossRef]
  61. Liu, L.; Cao, L.L.; Shao, Y.; Ménard, G.; Stephan, D.W. A Radical Mechanism for Frustrated Lewis Pair Reactivity. Chem 2017, 3, 259–267. [Google Scholar] [CrossRef]
  62. Rokob, T.A.; Hamza, A.; Stirling, A.; Soós, T.; Pápai, I. Turning Frustration into Bond Activation: A Theoretical Mechanistic Study on Heterolytic Hydrogen Splitting by Frustrated Lewis Pairs. Angew. Chem. Int. Ed. 2008, 47, 2435–2438. [Google Scholar] [CrossRef] [PubMed]
  63. Grimme, S.; Kruse, H.; Goerigk, L.; Erker, G. The mechanism of dihydrogen activation by frustrated Lewis pairs revisited. Angew. Chem. Int. Ed. 2010, 49, 1402–1405. [Google Scholar] [CrossRef] [PubMed]
  64. Rokob, T.A.; Bako, I.; Stirling, A.; Hamza, A.; Papai, I. Reactivity models of hydrogen activation by frustrated Lewis pairs: Synergistic electron transfers or polarization by electric field? J. Am. Chem. Soc. 2013, 135, 4425–4437. [Google Scholar] [CrossRef]
  65. Biswas, S.; Pramanik, A.; Sarkar, P. Computational studies on the reactivity of alkyl halides over (Al2O3)n nanoclusters: An approach towards room temperature dehydrohalogenation. Nanoscale 2016, 8, 10205–10218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Wischert, R.; Laurent, P.; Coperet, C.; Delbecq, F.; Sautet, P. γ-Alumina: The essential and unexpected role of water for the structure, stability, and reactivity of “defect” sites. J. Am. Chem. Soc. 2012, 134, 14430–14449. [Google Scholar] [CrossRef]
  67. Ghuman, K.K.; Wood, T.E.; Hoch, L.B.; Mims, C.A.; Ozin, G.A.; Singh, C.V. Illuminating CO2 reduction on frustrated Lewis pair surfaces: Investigating the role of surface hydroxides and oxygen vacancies on nanocrystalline In2O3−x(OH)y. Phys. Chem. Chem. Phys. 2015, 17, 14623–14635. [Google Scholar] [CrossRef]
  68. Bernal, S.; Calvino, J.J.; Gatica, J.; López Cartes, C.; Pintado, J.M.; Trovarelli, A. Catalysis by ceria and related materials. Catal. Sci. Ser. 2002, 2, 85–168. [Google Scholar]
  69. Carrasco, J.; Vilé, G.; Fernández-Torre, D.; Pérez, R.; Pérez-Ramírez, J.; Ganduglia-Pirovano, M.V. Molecular-level understanding of CeO2 as a catalyst for partial alkyne hydrogenation. J. Phys. Chem. C 2014, 118, 5352–5360. [Google Scholar] [CrossRef] [Green Version]
  70. Vilé, G.; Colussi, S.; Krumeich, F.; Trovarelli, A.; Pérez-Ramírez, J. Opposite Face Sensitivity of CeO2 in Hydrogenation and Oxidation Catalysis. Angew. Chem. Int. Ed. 2014, 53, 12069–12072. [Google Scholar] [CrossRef]
  71. Zhou, S.; Wan, Q.; Lin, S.; Guo, H. Acetylene hydrogenation catalyzed by bare and Ni doped CeO2(110): The role of frustrated Lewis pairs. Phys. Chem. Chem. Phys. 2022, 24, 11295–11304. [Google Scholar] [CrossRef]
  72. García-Melchor, M.; López, N. Homolytic products from heterolytic paths in H2 dissociation on metal oxides: The example of CeO2. J. Phys. Chem. C 2014, 118, 10921–10926. [Google Scholar] [CrossRef] [Green Version]
  73. Zhang, S.; Huang, Z.Q.; Ma, Y.; Gao, W.; Li, J.; Cao, F.; Li, L.; Chang, C.R.; Qu, Y. Solid frustrated-Lewis-pair catalysts constructed by regulations on surface defects of porous nanorods of CeO2. Nat. Commun. 2017, 8, 15266. [Google Scholar] [CrossRef]
  74. Zhou, S.; Gao, L.; Wei, F.; Lin, S.; Guo, H. On the mechanism of alkyne hydrogenation catalyzed by Ga-doped ceria. J. Catal. 2019, 375, 410–418. [Google Scholar] [CrossRef]
  75. Moon, J.; Cheng, Y.; Daemen, L.L.; Li, M.; Polo-Garzon, F.; Ramirez-Cuesta, A.J.; Wu, Z. Discriminating the role of surface hydride and hydroxyl for acetylene semihydrogenation over ceria through in situ neutron and infrared spectroscopy. ACS Catal. 2020, 10, 5278–5287. [Google Scholar] [CrossRef]
  76. Riley, C.; De La Riva, A.; Zhou, S.; Wan, Q.; Peterson, E.; Artyushkova, K.; Farahani, M.D.; Friedrich, H.B.; Burkemper, L.; Atudorei, N.-V.; et al. Synthesis of nickel-doped ceria catalysts for selective acetylene hydrogenation. ChemCatChem 2019, 11, 1526–1533. [Google Scholar] [CrossRef]
  77. Zhao, Y.; Jalal, A.; Uzun, A. Interplay between Copper Nanoparticle Size and Oxygen Vacancy on Mg-Doped Ceria Controls Partial Hydrogenation Performance and Stability. ACS Catal. 2021, 11, 8116–8131. [Google Scholar] [CrossRef]
  78. Shi, X.; Lin, Y.; Huang, L.; Sun, Z.; Yang, Y.; Zhou, X.; Vovk, E.; Liu, X.; Huang, X.; Sun, M.; et al. Copper Catalysts in Semihydrogenation of Acetylene: From Single Atoms to Nanoparticles. ACS Catal. 2020, 10, 3495–3504. [Google Scholar] [CrossRef]
  79. Huang, Z.-Q.; Zhang, T.; Chang, C.-R.; Li, J. Dynamic Frustrated Lewis Pairs on Ceria for Direct Nonoxidative Coupling of Methane. ACS Catal. 2019, 9, 5523–5536. [Google Scholar] [CrossRef]
  80. Wan, Q.; Chen, Y.; Zhou, S.; Lin, J.; Lin, S. Selective hydrogenation of acetylene to ethylene on anatase TiO2 through first-principles studies. J. Mater. Chem. A 2021, 9, 14064–14073. [Google Scholar] [CrossRef]
  81. Ghoussoub, M.; Yadav, S.; Ghuman, K.K.; Ozin, G.A.; Singh, C.V. Metadynamics-Biased ab Initio Molecular Dynamics Study of Heterogeneous CO2 Reduction via Surface Frustrated Lewis Pairs. ACS Catal. 2016, 6, 7109–7117. [Google Scholar] [CrossRef]
  82. Negreiros, F.R.; Camellone, M.F.; Fabris, S. Effects of thermal fluctuations on the hydroxylation and reduction of ceria surfaces by molecular H2. J. Phys. Chem. C 2015, 119, 21567–21573. [Google Scholar] [CrossRef]
  83. Jian, M.; Liu, J.-X.; Li, W.-X. Hydroxyl improving the activity, selectivity and stability of supported Ni single atoms for selective semi-hydrogenation. Chem. Sci. 2021, 12, 10290–10298. [Google Scholar] [CrossRef] [PubMed]
  84. Ye, J.; Johnson, J.K. Design of Lewis Pair-Functionalized Metal Organic Frameworks for CO2 Hydrogenation. ACS Catal. 2015, 5, 2921–2928. [Google Scholar] [CrossRef]
  85. Zhao, S.; Lin, L.; Huang, W.; Zhang, R.; Wang, D.; Mu, R.; Fu, Q.; Bao, X. Design of Lewis Pairs via Interface Engineering of Oxide-Metal Composite Catalyst for Water Activation. J. Phys. Chem. Lett. 2021, 12, 1443–1452. [Google Scholar] [CrossRef]
  86. Primo, A.; Neatu, F.; Florea, M.; Parvulescu, V.; Garcia, H. Graphenes in the absence of metals as carbocatalysts for selective acetylene hydrogenation and alkene hydrogenation. Nat. Commun. 2014, 5, 5291. [Google Scholar] [CrossRef] [Green Version]
  87. Sun, X.; Li, B.; Liu, T.; Song, J.; Su, D.S. Designing graphene as a new frustrated Lewis pair catalyst for hydrogen activation by co-doping. Phys. Chem. Chem. Phys. 2016, 18, 11120–11124. [Google Scholar] [CrossRef]
  88. Chen, Z.; Zhao, J.; Jiao, Y.; Wang, T.; Yin, L. Achieving efficient N2 electrochemical reduction by stabilizing the N2H* intermediate with the frustrated Lewis pairs. J. Energy Chem. 2022, 66, 628–634. [Google Scholar] [CrossRef]
  89. Hao, Y.-C.; Guo, Y.; Chen, L.-W.; Shu, M.; Wang, X.-Y.; Bu, T.-A.; Gao, W.-Y.; Zhang, N.; Su, X.; Feng, X.; et al. Promoting nitrogen electroreduction to ammonia with bismuth nanocrystals and potassium cations in water. Nat. Catal. 2019, 2, 448–456. [Google Scholar] [CrossRef]
  90. Zhang, Y.; Mo, Y.; Cao, Z. Rational Design of Main Group Metal-Embedded Nitrogen-Doped Carbon Materials as Frustrated Lewis Pair Catalysts for CO2 Hydrogenation to Formic Acid. ACS Appl. Mater. Interfaces 2022, 14, 1002–1014. [Google Scholar] [CrossRef]
Figure 1. (a) Structure of the first reported homogeneous FLP. (b) Scheme of the common strategy for constructing homogeneous FLP catalysts. (a) is reproduced from reference [24] with permission.
Figure 1. (a) Structure of the first reported homogeneous FLP. (b) Scheme of the common strategy for constructing homogeneous FLP catalysts. (a) is reproduced from reference [24] with permission.
Molecules 27 03734 g001
Figure 2. Role of (a) frustration and (b) electron field in mechanisms of H2 activation catalyzed by FLPs proposed by Rokob et al. and Grimme et al., respectively. Schemes of FLP-catalyzed H2 dissociation via (c) the ET mechanism (symbol * indicates antibonding orbital) and (d) the EF mechanism. (a), (b), and (c,d) are from references [62,63,64] respectively, with permission from John Wiley and Sons and American Chemical Society.
Figure 2. Role of (a) frustration and (b) electron field in mechanisms of H2 activation catalyzed by FLPs proposed by Rokob et al. and Grimme et al., respectively. Schemes of FLP-catalyzed H2 dissociation via (c) the ET mechanism (symbol * indicates antibonding orbital) and (d) the EF mechanism. (a), (b), and (c,d) are from references [62,63,64] respectively, with permission from John Wiley and Sons and American Chemical Society.
Molecules 27 03734 g002
Figure 3. Configurations of the transition states for CH4 dissociation on various surface Al–O pairs for surfaces at different OH coverages (0, 3 and 6 OH/nm2). Adapted from reference [66] with permission.
Figure 3. Configurations of the transition states for CH4 dissociation on various surface Al–O pairs for surfaces at different OH coverages (0, 3 and 6 OH/nm2). Adapted from reference [66] with permission.
Molecules 27 03734 g003
Figure 4. (a) Calculated energy profiles of the FLP (magenta line) and CLP (black line) catalyzed H2 dissociation of CeO2(110) with (magenta line) and without (black line) O vacancies, based on PBE via VASP. Color code: yellow, Ce; pink, O; white, H. (b) CDD (upper) and ELF (bottom) maps of H2 activation TS on CeO2(110). (c) Top and side views of Ni(left)/Ga(right)-doped CeO2(111). FLPs are marked by solid circles. Color code: yellow, Ce; red, O; blue, Ni; green, Ga. (d) Energy profiles of acetylene hydrogenation catalyzed by Ga/Ni-doped CeO2(111) and defect CeO2(111), based on PBE-D3 via VASP. (ad) were adapted from references [45,46,73,74], all with permission.
Figure 4. (a) Calculated energy profiles of the FLP (magenta line) and CLP (black line) catalyzed H2 dissociation of CeO2(110) with (magenta line) and without (black line) O vacancies, based on PBE via VASP. Color code: yellow, Ce; pink, O; white, H. (b) CDD (upper) and ELF (bottom) maps of H2 activation TS on CeO2(110). (c) Top and side views of Ni(left)/Ga(right)-doped CeO2(111). FLPs are marked by solid circles. Color code: yellow, Ce; red, O; blue, Ni; green, Ga. (d) Energy profiles of acetylene hydrogenation catalyzed by Ga/Ni-doped CeO2(111) and defect CeO2(111), based on PBE-D3 via VASP. (ad) were adapted from references [45,46,73,74], all with permission.
Molecules 27 03734 g004
Figure 5. (a) Potential FLPs assembled by Ti and O on anatase TiO2(101) and TiO2(001) with Ovs (denoted by blue circles). (b) Correlations between the integration of crystal orbital Hamilton population (ICOHP) of O–H with H2 activation energy, based on PBE-D3 via VASP. Adapted from reference [80] with permission.
Figure 5. (a) Potential FLPs assembled by Ti and O on anatase TiO2(101) and TiO2(001) with Ovs (denoted by blue circles). (b) Correlations between the integration of crystal orbital Hamilton population (ICOHP) of O–H with H2 activation energy, based on PBE-D3 via VASP. Adapted from reference [80] with permission.
Molecules 27 03734 g005
Figure 6. Collective variables (CVs) used in metaD AIMD simulations of the free energy surfaces of (a) H2 dissociation and (b) CO2 reduction catalyzed by In/O FLP at 20 and 180 °C. Color scheme: catalyst surface, white; LA (In), light blue; H, blue; O, red. Adapted from reference [81] with permission.
Figure 6. Collective variables (CVs) used in metaD AIMD simulations of the free energy surfaces of (a) H2 dissociation and (b) CO2 reduction catalyzed by In/O FLP at 20 and 180 °C. Color scheme: catalyst surface, white; LA (In), light blue; H, blue; O, red. Adapted from reference [81] with permission.
Molecules 27 03734 g006
Figure 7. (a) Charge density difference map of the transition state of H2 activation catalyzed by the Au/NH3 FLP. (b) Schematic illustration of the favorable interactions among Au, H2, and NH3, symbol * indicated the antibonding orbital. Adapted from reference [50] with permission.
Figure 7. (a) Charge density difference map of the transition state of H2 activation catalyzed by the Au/NH3 FLP. (b) Schematic illustration of the favorable interactions among Au, H2, and NH3, symbol * indicated the antibonding orbital. Adapted from reference [50] with permission.
Molecules 27 03734 g007
Figure 8. (a) Structures and adsorption of H2O on Ni13O13/Au(111), Ni10/Au(111), Ni10O6/Au(111), and Ni10O6-Ni6/Au(111). Color code: Au, yellow; metallic Ni, grey; oxidized Ni, blue; oxygen in NiOX, red; oxygen in H2O, brown; hydrogen, white. (b) Gibbs free energy diagram of H2O dissociation on the four surfaces (* indicated the adsorption site), based on PBE with vdW-DF, via VASP. (c) Electrostatic potential colored electron density of Ni1 0O6-Ni6/Au(111). Adapted from reference [85] with permission.
Figure 8. (a) Structures and adsorption of H2O on Ni13O13/Au(111), Ni10/Au(111), Ni10O6/Au(111), and Ni10O6-Ni6/Au(111). Color code: Au, yellow; metallic Ni, grey; oxidized Ni, blue; oxygen in NiOX, red; oxygen in H2O, brown; hydrogen, white. (b) Gibbs free energy diagram of H2O dissociation on the four surfaces (* indicated the adsorption site), based on PBE with vdW-DF, via VASP. (c) Electrostatic potential colored electron density of Ni1 0O6-Ni6/Au(111). Adapted from reference [85] with permission.
Molecules 27 03734 g008
Figure 9. (a) Structure and electron localized function (ELF) diagram of B doped phosphorene. (b) Energy diagram of H2 activation catalyzed by B doped phosphorene, symbol * indicated adsorption site here. Adapted from reference [54] with permission.
Figure 9. (a) Structure and electron localized function (ELF) diagram of B doped phosphorene. (b) Energy diagram of H2 activation catalyzed by B doped phosphorene, symbol * indicated adsorption site here. Adapted from reference [54] with permission.
Molecules 27 03734 g009
Figure 10. (a) Scheme of potential FLPs on Al-doped g-C3N4 (upper) and C2N (bottom). (b) Energy profile of H2 activation on Al/N FLPs. Based on PBE-D2 via VASP. (c) CDD (upper) and ELF (bottom) maps for TSs of H2 activation on g-C3N4-1 and C2N-1. Adapted from reference [55] with permission.
Figure 10. (a) Scheme of potential FLPs on Al-doped g-C3N4 (upper) and C2N (bottom). (b) Energy profile of H2 activation on Al/N FLPs. Based on PBE-D2 via VASP. (c) CDD (upper) and ELF (bottom) maps for TSs of H2 activation on g-C3N4-1 and C2N-1. Adapted from reference [55] with permission.
Molecules 27 03734 g010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wan, Q.; Lin, S.; Guo, H. Frustrated Lewis Pairs in Heterogeneous Catalysis: Theoretical Insights. Molecules 2022, 27, 3734. https://doi.org/10.3390/molecules27123734

AMA Style

Wan Q, Lin S, Guo H. Frustrated Lewis Pairs in Heterogeneous Catalysis: Theoretical Insights. Molecules. 2022; 27(12):3734. https://doi.org/10.3390/molecules27123734

Chicago/Turabian Style

Wan, Qiang, Sen Lin, and Hua Guo. 2022. "Frustrated Lewis Pairs in Heterogeneous Catalysis: Theoretical Insights" Molecules 27, no. 12: 3734. https://doi.org/10.3390/molecules27123734

Article Metrics

Back to TopTop