Experimental and computational aspects of molecular frustrated Lewis pairs for CO2 hydrogenation: en route for heterogeneous systems?

Riddhi Kumari Riddhi a, Francesc Penas-Hidalgo b, Hongmei Chen b, Elsje Alessandra Quadrelli a, Jérôme Canivet *a, Caroline Mellot-Draznieks *b and Albert Solé-Daura *cd
aIRCELYON, UMR 5256, Université LYON 1, 2 avenue Albert Einstein, 69626 Villeurbanne Cedex, France
bLaboratoire de Chimie des Processus Biologiques, CNRS UMR 8229, Collège de France, PSL Research University, Sorbonne Université, 75231 Paris Cedex 05, France. E-mail: caroline.mellot-draznieks@college-de-france.fr
cDepartment de Química Física i Inorgànica, Universitat Rovira i Virgili, Marcel·lí Domingo 1, Tarragona 43007, Spain
dInstitute of Chemical Research of Catalonia (ICIQ-CERCA), The Barcelona Institute of Science and Technology, Avgda. Països Catalans, 16, 43007 Tarragona, Spain. E-mail: asole@iciq.es

Received 22nd December 2023

First published on 30th August 2024


Abstract

Catalysis plays a crucial role in advancing sustainability. The unique reactivity of frustrated Lewis pairs (FLPs) is driving an ever-growing interest in the transition metal-free transformation of small molecules like CO2 into valuable products. In this area, there is a recent growing incentive to heterogenize molecular FLPs into porous solids, merging the benefits of homogeneous and heterogeneous catalysis – high activity, selectivity, and recyclability. Despite the progress, challenges remain in preventing deactivation, poisoning, and simplifying catalyst-product separation. This review explores the expanding field of FLPs in catalysis, covering existing molecular FLPs for CO2 hydrogenation and recent efforts to design heterogeneous porous systems from both experimental and theoretical perspectives. Section 2 discusses experimental examples of CO2 hydrogenation by molecular FLPs, starting with stoichiometric reactions and advancing to catalytic ones. It then examines attempts to immobilize FLPs in porous matrices, including siliceous solids, metal–organic frameworks (MOFs), covalent organic frameworks, and disordered polymers, highlighting current limitations and challenges. Section 3 then reviews computational studies on the mechanistic details of CO2 hydrogenation, focusing on H2 splitting and hydride/proton transfer steps, summarizing efforts to establish structure–activity relationships. It also covers the computational aspects on grafting FLPs inside MOFs. Finally, Section 4 summarizes the main design principles established so far, while addressing the complexities of translating computational approaches into the experimental realm, particularly in heterogeneous systems. This section underscores the need to strengthen the dialogue between theoretical and experimental approaches in this field.


image file: d3cs00267e-p1.tif

Riddhi Kumari Riddhi

Riddhi Kumari received her BSc from BHU, India and MSc from NIT Warangal, India. She did her Master thesis in LCC, Toulouse under the supervision of Prof. J.-B. Sortais and Dr S. Bastin. Currently, she is a PhD student under the direction of Dr Jérôme Canivet and Dr Alessandra Quadrelli at IRCELYON. She is working on the heterogenization of molecular catalysts into porous supports like metal–organic frameworks and porous organic polymers. A part of her thesis is dedicated to the heterogenization of FLPs in porous solids; and another part focuses on nickel coordination chemistry at bipyridyl sites within photoactive POPs.

image file: d3cs00267e-p2.tif

Francesc Penas-Hidalgo

Francesc Penas-Hidalgo received his BSc and MSc from Rovira i Virgili University (URV, Spain). He did his Master thesis under the supervision of Dr Albert Solé-Daura and Prof. Josep M. Poblet. Currently, he is pursuing a PhD under the supervision of Dr Albert Solé-Daura and Dr Caroline Mellot-Draznieks at Collège de France, in Paris. He is working on the computational modelling of the reactivity of Frustrated Lewis Pairs (FLPs) towards H2 splitting and CO2 hydrogenation. Moreover, he is computationally exploring other CO2 reduction approaches promoted by polyoxometalates and metalloporphyrins.

image file: d3cs00267e-p3.tif

Hongmei Chen

Hongmei Chen obtained her PhD in 2021 from the University of Liverpool, joining Prof. Andrew Cooper's group for her PhD studies and working on covalent organic frameworks (COFs) for photocatalytic water splitting. She then started her postdoctoral research at Collège de France on the hydrogenation of CO2 using metal–organic frameworks (MOFs)-based catalysts.

image file: d3cs00267e-p4.tif

Elsje Alessandra Quadrelli

Alessandra Quadrelli is director of research in chemistry from the French National Centre for Scientific Research, CNRS, at the IRCELYON laboratories. Her research focuses on organometallic mechanisms for grafting catalytically active transition metals or for growing thin films by atomic layer deposition (ALD) on surfaces (Silica, MOFs and silicon wafer substrates). Her current main target reaction is CO2 reduction.

image file: d3cs00267e-p5.tif

Jérôme Canivet

Jérôme Canivet was appointed CNRS researcher at the IRCELYON in 2010. He works at developing innovative catalytic processes for sustainable fine chemicals and energy. His research topics range from C–C coupling to asymmetry, photocatalysis and green fuels. In 2018, he received the Young Investigator Award from the Catalysis Division of the French Chemical Society for creating trends reducing the gap between homogeneous and heterogeneous catalysis. He further aims at exploiting the confinement of molecular catalytic systems into porous structures for the improvement of their catalytic activity and selectivity, and he is coordinating cooperative projects on this topic.

image file: d3cs00267e-p6.tif

Caroline Mellot-Draznieks

Caroline Mellot-Draznieks joined the CNRS in 1998 at Institut Lavoisier in Versailles. She then moved to the United Kingdom in 2005, where she held an Advanced EPRSC fellowship at UCL. She is now Director of Research (CNRS) at the Collège de France in Paris. Her research interests are related to nanoporous materials for adsorption and catalysis combining computational chemistry and experimental approaches, with a current major focus on CO2 conversion.

image file: d3cs00267e-p7.tif

Albert Solé-Daura

Albert Solé-Daura earned his PhD from Rovira i Virgili University (URV, Spain) in 2020. Subsequently, he conducted postdoctoral research at the Collège de France (France), the University of Amsterdam (The Netherlands), the URV as a Margarita Salas fellow, and the Institut Català d’Investigació Química (ICIQ, Spain). Recently, in 2024, he was awarded a “la Caixa” Junior Leader fellowship to initiate his independent research career at ICIQ. His research interests concern the application of computational methods to delve into the mechanisms of catalytic processes, primarily involving by polyoxometalates and Metal–Organic Frameworks.


1. Introduction

Catalysis plays a key role in the manufacturing processes underlying our chemical industry. Still, there is an urgent need to reduce its environmental impact and to move towards a more sustainable society, which requires a careful reassessment of certain fundamental aspects. Many homogeneous and heterogeneous catalytic methods rely on transition metals as active sites. The use of such potentially scarce resources is coupled with limited recyclability and consequential resource waste in homogeneous systems, or with metal recovery using energy-intensive processes in heterogeneous systems.1 Simultaneously, the direct conversion of small molecules into valuable synthetic intermediates, particularly employing CO2 as a readily available and cost-effective C1 reservoir,2 gains importance in increasing the sustainability of associated chemical processes.3 Therefore, there is a persistent requirement to develop novel catalytic systems based on readily available, cost-effective elements with wide-ranging applicability, producing high-value compounds for environmentally friendly chemical production.

The advent of molecular frustrated Lewis pairs (FLPs), introduced by Stephan et al. in 2006,4 showcases reactivity stemming from steric hindrance between Lewis acid (LA) and base (LB) partners, precluding the formation of the LA–LB adduct.5–7 FLPs, often referred to as a “catalysis Holy Grail”, have ushered in a new era of sustainable green chemistry. In particular, they allow the transition metal-free activation of small molecules, demonstrating significant potential in hydrogen activation and double bond reduction in fine chemicals, including CO2 hydrogenation.8–11

While the field of homogeneous FLPs has gathered substantial knowledge in the past decade,12,13 the transfer of FLPs’ chemistry to heterogeneous catalysis is rather recent and garnering increasing interest. Besides addressing recyclability issues, the heterogenization of molecular FLPs allows the isolation of active sites within a controlled chemical environment, while preventing undesired interactions and catalyst deactivation. This isolation alters substrate activation interactions on the heterogeneous catalyst, leading to unprecedented reactivities explicable through molecular chemistry mechanisms. The development of heterogeneous FLPs is currently challenging and mainly limited to doped/defective oxide and metal surfaces.14,15

Incentives for discovering heterogeneous FLPs encompass aspects ranging from molecular-scale control over the distance between FLP's partners, influencing reactivity, to their recyclability, shaping, and scale-up production. Quantum mechanical calculations serve as a powerful tool for the in silico design of molecular architectures, having matured in molecular catalysis for predicting efficient systems, remarkably for asymmetric synthesis.16–18 However, computational methodologies are at an earlier stage concerning heterogeneous catalysis.19–22

In addition to the perspectives on FLPs’ chemistry by Stephan et al.,14,23 recent reviews in the last five years have covered distinct advancements in FLPs’ chemistry by considering either the realm of homogeneous catalysis24–26 or strategies for their heterogenization into solid materials.15,27,28 A recent review covered the specific area of the computational chemistry of FLPs, focusing exclusively on molecular systems.29 Therefore, there remains a critical need for a comprehensive analysis of the recent progress and existing challenges in the discovery of FLP-based catalytic systems for CO2 hydrogenation, encompassing both molecular and solid-state FLPs and integrating recent experimental and theoretical facets.

This review delves into fifteen years of FLPs chemistry development, exploring the chemistry of FLPs based on main group elements from experimental discovery to computational mechanistic dissection. The discussion encompasses the recent experimental advances of FLPs in catalytic processes, tools developed for their heterogenization, and key aspects of structure–activity relationships for CO2 hydrogenation as determined in silico.

2. Experimentally reported FLPs for CO2 hydrogenation

Since the pioneering report by Stephan et al. in 2006 with H2 splitting by a transition metal-free P/B containing molecule,4 this field exhibits an ever-expanding range of molecular LB (B, C, N, O, S, Se and Te) and LA (B, Al, Ga, In, C, Si, Sn, N and P) partners in intra- or inter-molecular fashions.9,30 They demonstrated their large potential for the activation of H2 and the subsequent reduction of double bonds in fine chemicals,8,25 including CO2 hydrogenation.31–33 We will first summarize here the main experimental findings on FLP-mediated CO2 hydrogenation using molecular H2 (Table 1).
Table 1 Overview of molecular FLP systems for CO2 reduction. Data were extracted from references when reported. (RT: room temperature)
FLP formula LA/LB pair Inter/intramolecular Solvent T, P, time Product (yield) Ref.
B(C6F5)3/TMP B/N Inter Toluene 160 °C, 1 eq. CO2, excess H2, 6 days MeOB(C6F5)2 but CH3OH after distillation (17–25%) 31
B(C6F5)3/lutidine B/N Inter Toluene RT, 4 atm H2 and CO2 [LutH] [HC([double bond, length as m-dash]O)OB(C6F5)3] 34
B(C6F5)3/amine B/N Inter Bromobenzene RT, 4 atm H2 [PhNMe2H][HB(C6F5)3] 35
B(C6F5)3/tBu3P B/P Inter Bromobenzene 25 °C, 1 bar CO2, immediate tBu3P(CO2)B(C6F5)3 (87%) 36
(Me3C6H2)2PCH2CH2B(C6F5) B/P Intra Pentane 25 °C, 2 bar CO2, 15 min Cyclo-(Me3C6H2)2PCH2CH2–B(C6F5)2–(CO2) (79%)
(C6F4(C6F5))3B/P(tBu3) B/P Inter Toluene 145 °C, 1 atm CO2, 1 atm H2, 24 h [HPt-Bu3][(C6F4(C6F5))3BO2CH] 37
AlX3/Mes3P Al/P Inter Bromobenzene 25 °C, 1 atm CO2, 2 h Methanol (37–51%) 38
AlX3/Mes3P Al/P Inter Bromobenzene RT, 2 atm CO2, 16 h CO, Mes3P(C(OAlI2)2O)(AlI3) and [MesPX][AlX4] 39
[MeC(CH2PPh2)3Cu(NCMe)][PF6]/C10H16N2 (DBU) Cu/N Inter Acetonitrile 80–140 °C, 40 atm H2/CO2, 6–20 h Formate salt[C10H17N2][HCO2] 40
1-BR2-2-NMe2-C6H4 B/N Intra Benzene 80 °C, 1 atm CO2, 4 atm H2, 9 days Formates, acetals, methoxides 32
tBu2P–O–Al Al/P Intra n-Hexane −196 °C then RT, 2 eq. CO2, overnight tBu2P(H)–O–Al(CO2H) 41
K2[(BCF)2–CO3] B/CO32− Inter THF 160 °C, 40 bars H2, 20 bars CO2, 48 h BCF-HCO2K 42
Tris(bromo)tridurylborane/DBU B/N Inter Acetonitrile 120 °C, 130 bar, 48 h Formate 43


2.1. Molecular FLPs reported for non-catalytic and catalytic CO2 reduction

In this section, we aim to discuss the development of molecular FLPs and the study of their reactivity in biphasic gas–liquid systems from CO2 activation to CO2 hydrogenation, leading to unique catalytic systems. Here, FLPs are classified as either structurally intramolecular, i.e. the Lewis base (LB) and Lewis acid (LA) are part of the same molecule, or intermolecular, i.e. the LB and LA partners belong to two independent molecules.
2.1.1. Inter- and intramolecular FLPs based on BCF as Lewis acid. Tris(pentafluorophenyl)borane (BCF, B(C6F5)3) is a widely used super Lewis acid which was first synthesised in the 1960s by Massey and coworkers44,45 and used as an activator or cocatalyst for alkenes polymerization reactions.46–48

In 2009, Mömming, Stephan and coworkers described for the first time a novel approach for the metal-free binding of CO2 by FLPs based on the combination of BCF and phosphine.36 They demonstrated that the tBu3P/B(C6F5)3 FLP pair dissolved in bromobenzene can react with CO2 under ambient pressure at room temperature to yield a white solid product tBu3P(CO2)B(C6F5)3 in high yield (Fig. 1a). Specifically, the LB component (tBu3P) interacts with the carbon atom, while the LA component (BCF) stabilizes an oxygen atom of CO2 via coordination. They also showed that a pentane solution of an intramolecular FLP, (Me3C6H2)2PCH2CH2B(C6F5)2, could bind to CO2 at room temperature to form a white solid cyclo-(Me3C6H2)2PCH2CH2B(C6F5)2–(CO2) with 79% yield (Fig. 1b). The reversible nature of the CO2 binding was demonstrated upon heating the FLP–CO2 adducts under vacuum to 80 °C for 5 hours. Under these conditions, about half of the carbon dioxide was released and the initial FLP mixture was regenerated. These results were further supported by DFT calculations. This pioneering work offered a new metal-free approach to the activation of CO2, which avoids the drawbacks associated with traditional metal-based catalysts. Also, it provided unprecedented insights into the fundamental chemistry of CO2 binding by FLPs, which inspired further research in this area and led to potential applications in CO2 capture and recycling.


image file: d3cs00267e-f1.tif
Fig. 1 CO2 binding on P/BCF-based FLP.

In the same year 2009, Ashley et al. illustrated the first homogenous selective hydrogenation of CO2 to CH3OH using an FLP-based non-metal-mediated procedure at low pressures (Fig. 2).31 The FLP mixture used here was a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 mixture of 2,2,6,6- tetramethylpiperidine (TMP, Me4C5NH) and BCF which was already proven to cleave H2 heterolytically giving a salt [TMPH][HB(C6F5)3] at low pressure (1–2 atm).49 They discovered that this FLP mixture could not only cleave H2 but also subsequently insert CO2 in the B–H bond.


image file: d3cs00267e-f2.tif
Fig. 2 Multistep FLP-mediated CO2 hydrogenation.

Other research groups have also employed a similar two-step mechanism, as described above, in their studies.50,51 The vacuum distillation of the solvated methoxyborate species resulted in the release of CH3OH (17–25% yield) as the only selective C1 product, along with C6F5H and TMP. The only source of labile protons in the decomposition reactions is provided by [TMPH]+, which triggers ion recombination to form TMP and CH3OH·B(C6F5)3. Dissociation of this hydroxy adduct could lead to the FLP's regeneration in a catalytic system. However, at 160 °C, the H+ attack on the ipso-C proceeds faster, resulting in Lewis acid breakdown. As a result, inert boroxine CH3OB(C6F5)2 is finally formed. The identification of these key intermediates along the reaction pathway was made possible through mechanistic investigations and multinuclear solution NMR spectroscopy.

Geier and Stephan reported that the FLP based on 2,6-lutidine (Lut = 2,6-dimethylpyridine) and B(C6F5)3 could heterolytically cleave H2 to produce the borohydride salt, [LutH][HB(C6F5)3], as a white solid (Fig. 3).52 Later, Mayer et al. explored the interaction of CO2 with the same salt.34 By treating it with 4 atm of CO2 in toluene-d,8 they observed the formation of an air-stable white solid, [LutH][HC([double bond, length as m-dash]O)OB(C6F5)3]. The reduction of CO2 by [LutH][HB(C6F5)3] occurs much faster than the corresponding reaction with TMP, despite computational results by Pápai et al. suggesting that TMP/B(C6F5)3 is a more reactive FLP.53 The calculations showed that heterolytic cleavage of H2 is thermodynamically less favorable for Lut/B(C6F5)3 than for TMP/B(C6F5)3 by approximately 10 kcal mol−1. It was suggested that the higher reactivity of Lut/B(C6F5)3 for CO2 reduction may be due to a mechanism involving both hydride donation from [HB(C6F5)3] and activation of CO2 by free B(C6F5)3. However, preliminary studies have shown that adding excess B(C6F5)3 did not significantly accelerate the reaction, and adding excess lutidine did not cause any substantial change in rate.


image file: d3cs00267e-f3.tif
Fig. 3 Reduction of CO2 by Lut/B(C6F5)3 to the formato complex [LutH][HC([double bond, length as m-dash]O)OB(C6F5)3].

Later, Voss and coworkers examined how different amines (N-dimethylaniline, N-isopropylaniline, 1,4-C6H4(CH2NHtBu)2, and benzyldimethylamine) reacted with BCF and explored their FLP behavior.35 They illustrated how both steric and electronic factors may influence the formation of LP adducts. Although these factors may hinder adduct formation, they allow access to free FLPs that can be used for further reactions. Additionally, the borane's hydridophilic nature promotes the formation of minor side products that are derived from hydride abstraction from the α-C to produce iminium salts. They further studied the reaction of H2 and CO2 with these pairs. The Lewis pair of N-dimethylaniline–B(C6F5)3 reacted quickly with 4 atm H2 at 25 °C, even though the LB component had low basicity. By the heterolytic cleavage of H2, it produced the ammonium–hydridoborate salt [PhNMe2H][HB(C6F5)3]. The CO2 adducts of such salts have been investigated by Erker et al. as discussed above and were found to be thermally unstable, releasing CO2 quickly in solution at temperatures ranging from −20 to +80 °C.36,54

The following year, Travis et al. focused specifically on FLPs based on the Lewis acids tris(perchloroaryl)borane (BArCl) and tris(2,2′,2′′-perfluorobiphenyl)borane (PBB) with selected phosphines, which were extensively studied for small molecules activation.37,55,56 The reactivity was found mainly driven by the activation of H2 before a concerted addition of CO2. The Lewis acidity of the catalyst is crucial in this process, as it needs to be strong enough to split H2 but not too strong to allow hydride transfer to CO2. They designed Lewis acid assuming that the steric bulkiness at the ortho position could decrease the B–O bond strength of the methoxyborate formed and could consequently facilitate its cleavage to evolve methanol and regenerate the FLP catalyst. Such sterically congested LA was found less effective in activating H2 despite having higher Lewis acidity. This was mainly ascribed to the steric hindrance around the boron center, hindering the approach of H2 and increasing the energy barrier for the reaction.

Among all the phosphines tested with BArCl, only P(tBu)3 and P(Cy)3 could cleave H2 without thermal decomposition and were found to be the most active ones, completing the reaction in a long time of 56 and 40 hours at 90 °C, respectively. The formatoborate [PBB-OC(O)H][H-P(tBu)3] was prepared by exposing the [PBB-H][H-P(tBu)3] salt to one atmosphere of CO2. Still, the reaction did not occur at room temperature but required heating at 140 °C for several days.

2.1.2. Intermolecular FLPs based on other metal-based Lewis acids. In contrast to previous reversible systems, Menard and Stephan reported in 2010 FLPs based on aluminium AlX3 (X = Cl or Br) with PMes3 (Mes = 2,4,6-C6H2Me3) which could irreversibly bind CO2. This system could further produce methanol at room temperature in the presence of excess ammonia borane as a hydrogen source (Fig. 4).38 The CO2 sequestered species Mes3PC(OAlX3)2 (X = Cl or Br) was stable and obtained with yields up to 83%. The corresponding CO2 adducts were stable up to 80 °C under vacuum, despite having weaker Lewis acidic and basic components than other similar compounds that rapidly lose CO2 at much lower temperatures.36 Quenching with H2O led to the isolation of CH3OH in 37–51% yield, representing a further example of the stoichiometric reduction of CO2 to CH3OH. However, no mechanism of this reduction reaction was reported. Since then, and over the last decade, Stephan and co-workers widely studied various FLPs for CO2 activation and/or reduction following a similar methodology.8,57–60
image file: d3cs00267e-f4.tif
Fig. 4 Reactivity of Al/P FLP towards CO2 and subsequent reduction reaction.

The same group demonstrated that P/Al/CO2 adducts can also be used as a means to reduce CO2 to CO.39 Berke et al. used transition metal-based FLPs in the hydrogenation of CO2 in 2013.61 The authors showed that 0.5 mol% of ReHBr(NO)(PR3)2 and B(C6F5)3 in the presence of C5H6Me4NH catalyzed the formation of formate salt [C5H6Me4NH2][HCO2]. Similarly, in 2015, Zall et al. reported the CO2 hydrogenation catalyst based on a well-defined copper complex with triphos ligand 1,1,1-tris-(diphenylphosphinomethyl)ethane.40 The [MeC(CH2PPh2)3Cu (NCMe)][PF6] complex acted as a Lewis acid in the presence of C10H16N2 (DBU) base, H2, and CO2 to generate the salt [C10H17N2][HCO2].

2.1.3. Intramolecular FLPs based on Al or B Lewis acid. After reporting the reduction of CO2 using amphiphilic FLPs in the presence of hydroboranes giving methoxyboranes,62–65 Fontaine et al described in 2015 intramolecular FLPs containing B and N as Lewis pairs.32 Instead of using BCF with high Lewis acidity, they explored less acidic systems containing tri-aryl boron centers. Like Menard and Stephan in 2010 for intermolecular systems using Al/P FLP pairs,38 Mitzel et al. recently reported FLP based on Al/P pairs bridged by geminal oxygen.41 Bis2AlH (Bis = CH(SiMe3)2) and tBu2P(O)H were reacted to form an adduct of phosphane and alane, tBu2P(H)–O–Al(H)Bis2 (Fig. 5). Due to the high reactivity of the latter releasing H2, its reaction with CO2 was performed at −196 °C. This allowed achieving a stoichiometric reduction of CO2, resulting in the formation of stable aluminium formate. Despite its appealing reactivity, this FLP system could not be used as a catalyst for the hydrogenation of CO2 as the H2 adduct was found to be in equilibrium with free H2 and the CO2 adduct was too stable.
image file: d3cs00267e-f5.tif
Fig. 5 Reactivity of Al–O–P FLP towards H2 and CO2 to formic acid adduct.

More recently, Fernández, Breher and co-workers reported a “hidden FLP” based on a five-membered ring phosphorus ylide containing aluminium or gallium as a LA site.66 This system was shown to reduce CO2 to methanol in the presence of borane as a reducing agent and upon the addition of water.

2.1.4. Emerging catalytic molecular FLPs for CO2 hydrogenation. Towards the catalytic application of FLPs, a seminal work by Stephan and coworkers in 2014 reported the combination of P(tBu)3 as Lewis base in a catalytic amount with an excess of 9-borabicyclo(3.3.1)nonane (H-BBN) as Lewis acid. Based on borane consumption, TONs of 5500 were reached for the CO2 activation to methoxy–borane (MeO–BBN) with 98% selectivity in bromobenzene solvent under 5 atm. at 60 °C (Fig. 6).67
image file: d3cs00267e-f6.tif
Fig. 6 Catalytically active Lewis base in FLP-driven CO2 activation.

Later in 2018, Zhao et al. described a similar “semi-catalytic” CO2 hydrogenation to formate using a catalytic amount of BCF as LA and an excess of carbonate base M2CO3 (M = Na, K, and Cs).42 This system resulted in the formation of a Lewis pair (K2–[(BCF)2–CO3]) that could react with both H2 and CO2 to produce the BCF–HCO2M (Fig. 7). Using a 10[thin space (1/6-em)]000-fold excess of base compared to BCF, under CO2[thin space (1/6-em)]:[thin space (1/6-em)]H2 = 20[thin space (1/6-em)]:[thin space (1/6-em)]40 bar, in THF for 48 hours at 160 °C, they could reach a TON of 4000 based on the moles of formate produced per mole of BCF.


image file: d3cs00267e-f7.tif
Fig. 7 Proposed catalytic process for the hydrogenation of CO2 using K2-[(BCF)2–CO3]. Reproduced from ref. 42. Copyright 2019 Wiley-VCH.

More recently, in 2022, Dyson and Corminbœuf used linear scaling relationships to explore FLPs combinations for the direct hydrogenation of CO2.43 Thousands of FLPs combinations were computationally screened based on the acidity and basicity of the individual components. Their computational approach and results are discussed in more details in Section 3.2.2. In short, the authors showed that balancing the cumulative strength is key to catalytic performance. From the wide set of FLPs explored, those consisting of tris(p-bromo)tridurylborane (tbtb)/DBU with a weak LA and B-Me2F2/pyridine with a weak LB were found to be the most effective ones. Their predicted high activity results from their ability to provide a suitable balance between the limiting steps, i.e. activation of H2 and release of the product. Notably, the tbtb/DBU-based FLP was found to experimentally catalyze the direct hydrogenation of CO2 to formate. It occurs in the presence of 100 equivalents of the base compared to the acid at a very high pressure of 150 bar (25 bar CO2 and 125 bar H2) at 120 °C after 48 hours (Fig. 8). TON of 24 were obtained, and, noteworthy, the product of the reaction was the formate salt of the Lewis base. They also explained the reason why previous attempts to use molecular FLPs for CO2 hydrogenation combining strong LAs like BCF with various nitrogen-containing bases (such as TMP or lutidine) resulted in boroformate salts with no product release. The authors concluded that the formate release step would always be the limiting factor and prevent catalytic turnover, no matter which base is chosen when such a strong LA is used.


image file: d3cs00267e-f8.tif
Fig. 8 Catalytic hydrogenation of CO2 to formate using tbtb/DBU system. Adapted from ref. 43. Copyright 2022 Wiley-VCH.

2.2. FLPs heterogenization in porous solids for CO2 hydrogenation

Although FLP systems are renowned for their metal-free nature and great reactivity and selectivity, their immobilization into porous supports has been driven by the aim to enhance their stability against deactivation. Immobilization would also in principle allow better control of acid–base distances, a key parameter, especially for CO2 activation.68 Furthermore, the heterogenization of FLPs within porous solids should enable their implementation in continuous flow processes, also key for CO2 valorization at a large scale. While the field of homogeneous FLPs has benefited from the great knowledge gained over the last decade,12,13 the transfer of FLPs’ chemistry to heterogeneous catalysis is very recent and remains challenging.14,15 We gather here examples of molecular FLPs heterogenization on various supports for CO2 and/or H2 activation (Table 2).
Table 2 Overview of heterogenized FLP systems for CO2 and/or H2 activation. Data were extracted from references when reported. (RT: room temperature)
FLP type LA/LB pair Inter/Intramolecular Solvent T, P, time Product (yield) Ref.
HB(C6F5)2/PPh3@silica B/P Inter Pentane 80 °C, 40 bar H2, 4 h [[triple bond, length as m-dash](SiO)2Al(OEt)(OC6H4PH(C6H5)2)][HB(C6F5)2] 69
BCF@silica/tBu3P B/P Inter Toluene 65 °C, 2 bar H2, 72 h [[triple bond, length as m-dash]SiOB(H)(C6F5)2][tBu3PH] 70
PR3@silica/BCFPR3/BCF@silica B/P Inter Pentane, toluene RT to 60 °C, 2 bar H2, 2 bar CO2, 48 h Formic acid 71
B(C6F5)2(Mes)/DABCO@MOF B/N Inter Toluene RT, 10 bar H2, 48 h α,β-unsaturated amine 72
B(C6F5)3/DABCO@MOF B/N Inter Toluene 80 °C, 60 bar H2, 24 h Alkylidene malonate (88–95%) 73
B-MOF/amine B/N Inter Acetonitrile 120 °C, 10 bar CO2, 24h Benzimidazoles 74
B(C6F5)3/MOF(porphyrin) B/N Inter Toluene RT to 100 °C, 20 bars H2 and CO2, 21 h [(MeO)B(C6F5)3] then MeOH after hydrolysis 75
B(C6F5)3/diamine@MOF B/N Inter Toluene RT, 20 bar H2, 48 h Amine 76
BCF/P-POP B/P Inter Cyclohexane RT, 6 bar H2 [(Ar)3PH][HB(C6F5)3] 77
BCF/N-POP B/N Inter Toluene 80 °C, 60 bar H2, 24 h Diethyl-2-benzylmalonate 78
P-POP/B-POP B/P Inter Toluene 60 °C, 1 atm CO2 Formamide 79


2.2.1. Immobilization in siliceous porous solids. Taoufik et al. reported in 2016 a strategy to support triphenylphosphine and HB(C6F5)2 FLP on silica.69 The grafting of the phosphine was accomplished by reacting (4-hydroxyphenyl)biphenyl with ≡(SiO)2Al(iBu)(OEt) in ether at room temperature for 12 hours. The resulting silica-supported LB was exposed to the benzene solution of HB(C6F5)2 or BCF for 3 hours at room temperature to achieve the semi-immobilized FLP on silica support (Fig. 9). The same year, O’Hare reported a synthesis of silica-supported BCF ([triple bond, length as m-dash]SiO–B(C6F5)2, s-BCF) and then combined it with tri-tert-butylphosphine as LB to form a solid-phase [[triple bond, length as m-dash]SiO–B(C6F5)2][tBu3P].70 This FLP was shown to cleave H2 heterolytically in toluene at 65 °C for 72 hours to give [[triple bond, length as m-dash]SiO–B(H)(C6F5)2][tBu3PH]. Additionally, this FLP was shown to activate polar O–D bonds at room temperature. In 2018, the group of Zhang described the semi-heterogenization of Lewis pairs on silica.80 A silicon wafer functionalized with N-heterocyclic olefin as LB was shown to initiate lactones polymerization in the presence of tris(pentafluorophenyl)alumane as LA, giving linear polymer brushes. In 2020, Erasmus et al. reported a new approach to graft FLPs on hydroxylated silica nanopowder.71 They grafted triethoxysilane derivatives of Lewis acid and bases on the hydroxylated silica nanopowder which could interact with their respective LB and LA components to form FLPs. These silica nanopowder-supported LA and LB could activate H2 when exposed to their complementary FLP partners (2 bar, room temperature, 3 days) and subsequently transform CO2 into HCOOH (2 bar, 60 °C, 3 days) (Fig. 10).
image file: d3cs00267e-f9.tif
Fig. 9 Formation of silica-supported phosphine Lewis base combined with solubilized Lewis acids. Reprinted from ref. 69.

image file: d3cs00267e-f10.tif
Fig. 10 CO2 capture by silica-supported LA and dissolved LB mixture. Reprinted from ref. 71. Copyright 2020 American Chemical Society.
2.2.2. Immobilization in ordered metal–organic and covalent organic frameworks. Metal–organic frameworks (MOF) already demonstrated their appealing properties for CO2 capture and conversion.81–85 Thanks to their hybrid organic–inorganic nature, the tunability of their pore sizes, a large variety of topologies and their high crystallinity and high porosity,86,87 MOFs are ideal candidates for the heterogenization of FLPs. Indeed, crystalline-ordered networks, with regularly distributed functional groups, can allow precise control over the positioning of FLP partners, critically related to the FLP activity.

Since 2015, several computational studies have proposed the possibility of creating MOF-based FLP catalysts.88–93 However, only in 2018 the Ma's group achieved experimentally the introduction of FLPs into MOF in a semi-immobilized manner.73 The Cr-MIL-101 was selected as an adequate MOF platform due to its large open porosity and the presence of Cr open metal sites.94 The selected classical FLP comprises BCF as the LA and 1,4-diazabicycl[2.2.2]octane (DABCO) as a potent LB. One of the N atoms in DABCO is coordinated to the open metal site of the MOF and the second N atom is paired to BCF simultaneously as illustrated in Fig. 11. Although the obtained Cr-MIL-101-LP was not reported for CO2 hydrogenation, the solid efficiently catalyzed the reduction of C[double bond, length as m-dash]N imine bonds with up to 100% yield using HBpin as a reducing agent, as well as the direct hydrogenation of alkylidene malonates under 60 bars H2 for 24 hours at 80 °C giving up to 91% yield. This catalyst was recyclable for up to seven runs for the reduction of imines.


image file: d3cs00267e-f11.tif
Fig. 11 Synthesis of Cr-MIL-101-LP as a catalyst for C[double bond, length as m-dash]C and C[double bond, length as m-dash]N reduction reactions. Adapted from ref. 73. Copyright 2019, with permission of Elsevier.

The following year, in 2019, the same group reported the chemoselective catalytic hydrogenation of α,β-unsaturated imine compounds under 10 bar H2 at room temperature in high yields up to 100% using the same Cr-MIL-101 MOF.72 The FLP anchored in the MOF was B(C6F5)2(Mes)/DABCO following the same strategy.73 Interestingly, this immobilized FLP could activate H2 to form MIL-101(Cr)-FLP-H2 at room temperature and hence, form an ammonium hydridoborate salt, [CH(CH2CH2)3NH] [HB(C6F5)2(Mes)]. Powder XRD was used to confirm the retention of the structural integrity of MIL-101(Cr)-FLP-H2, its surface area being of 1120 m2g−1 as obtained by N2 sorption studies. The MIL-101(Cr)-FLP displayed the same productivity for the imine reduction for up to five runs.

Dyson and Stylianou and co-workers reported a water-stable MOF named SION-105 which incorporated a bulky LA-functionalized ligand.74 MOF SION-105 was synthesized by the combination of Eu-(NO3)3 and tris(p-carboxylic acid)tridurylborane (H3tctb) in a 2[thin space (1/6-em)]:[thin space (1/6-em)]1 mixture (Fig. 12). SION-105 allowed for the in situ formation of FLP by employing Lewis basic diamine substrates which could be efficiently transformed to benzimidazoles in the presence of excess silane and 10 bar CO2 as a C1 source at 120 °C for 24 hours, with isolated yields up to 90%.


image file: d3cs00267e-f12.tif
Fig. 12 Synthesis of SION-105 and the reaction of aromatic o-diamines with CO2 in the presence of SION-105. Reprinted with permission from ref. 74. Copyright 2020 Wiley-VCH.

In the same year, the same group conducted another study on stoichiometric CO2 fixation using an inverse approach.75 In the MOF-545 hydride framework, which contains Zr6O8 clusters linked by tetrakis(4-carboxyphenyl)porphyrin (TCPP) ligands, nitrogen atoms in porphyrin act as LB and formed an in situ FLP when combined with the BCF LA introduced as a guest molecule (Fig. 13).


image file: d3cs00267e-f13.tif
Fig. 13 BCF-impregnated MOF-545 for CO2 fixation.

The BCF@MOF-545 was subjected in toluene to 20 bars of H2 at room temperature for 1 hour and then to 20 bars of CO2 at 100 °C for approximately 20 hours. Following this procedure, the formation of methoxyborate, [(MeO)B(C6F5)3] was observed along with the side products methoxyborate and the hydroxyborate, [(HO)B(C6F5)3], which were confirmed by mass spectrometry and NMR. During the experiment, the decomposition of the products into [(MeO)B(C6F5)2] and C6F5H was observed. These products could potentially give methanol upon hydrolysis. Notably, the MOF host remained intact even after three runs. Although the NMR analysis showed the presence of the hydrogenated products, these were not quantified.

With a similar approach as in 2019, Ma's group developed a molecular chiral catalyst by incorporating chiral FLPs (CFLPs) into achiral MOFs for asymmetric hydrogenation of imines.76 Guided by a computational screening, the 2,5-dihydro-3,6-dimethoxy-2-isopropylpyrazine bifunctional basic amines were selected to coordinate the Cr(III) open site with one N atom and paired with BCF. The presence of the FLP was confirmed by elemental analysis, FTIR, XPS, EDS and 11B NMR comparison of CFLP@MIL-101(Cr)-H2 and BCF. When applied to the asymmetric hydrogenation of imines (20 bars of H2, 48 hours at room temperature) this catalyst gave yields above 95% in most of the different imines with enantiomeric excess up to 85%.

More recently, in 2022, a covalent organic framework (COF) functionalized with FLP was reported by Hu's group.95 Inspired by the catalytic activity of triaryl phosphine and BCF pair, they developed COF@FLP by a post-synthetic modification of the crystalline brominated COF with triaryl phosphine unit as Lewis base by cross-coupling reaction, followed by the encapsulation of BCF as Lewis acid. These COFs were used for the stereoselective hydrogenation of alkynes into Z-alkene with H2 and effectively recycled for up to ten runs.

2.2.3. Immobilization in disordered organic polymers. In 2017, the first microporous polymer networks based on tris(2-methylphenyl)phosphine and tris(2,6-dimethylphenyl)-phosphine were reported by Thomas et al. which contained the LB component of the FLP in its structure (Fig. 14).77 Polymers were shown to rapidly absorb BCF by swelling when suspended in a DCM solution of BCF to form a semi-heterogenized FLP. Fluorescence experiments under UV irradiation clearly showed the quenching of the fluorescence for the polymers due to the interaction with BCF. Solid-state 31P MAS NMR spectroscopy also confirmed the presence of P/B FLP. These polymers were proven capable of activating H2 at room temperature via H/D isotope scrambling experiments under a 6 bar H2/D2 mixture at room temperature.
image file: d3cs00267e-f14.tif
Fig. 14 BCF-impregnated phosphine-based polymer for H2 cleavage. Reprinted with permission of ref. 77. Copyright 2017, American Chemical Society.

Similarly, in 2018, Rose et al. reported a polyamine organic framework with tertiary amine as LB sites, which forms in situ FLP when exposed to BCF in solution.78 The impregnation of BCF was studied by ATR IR spectroscopy showing a new IR band at 1276 cm−1 corresponding to the B–N interaction between BCF and polyamine. As a semi-heterogenized FLP, this polymeric amine in combination with BCF was shown to cleave H2 heterolytically, as confirmed by 11B NMR, and applied in the catalytic hydrogenation of electron-poor diethyl benzylidenemalonate.

Following a strategy based on fully covalently immobilized FLP, Yan and coworkers reported the physical mixture of two linear polymers functionalized with either BCF derivative (LA) or triarylphosphine (LB) as illustrated in Fig. 15.79 Upon CO2 exposure under 1 atm. in toluene, the intermolecular polymeric FLP is assembled to form micelles with bridging CO2 between polymer chains. These micelles were shown to be nanocatalysts for the secondary amine formylation with TON up to 15000 in the presence of phenylsilane as a reducing agent, with efficient recycling for up to eight catalytic runs.


image file: d3cs00267e-f15.tif
Fig. 15 Self-assembled FLP–polymer micelles for CO2 activation. Reprinted with permission from ref. 79. Copyright 2018, WILEY-VCH.

From siliceous materials to MOFs, COFs and POPs, the above-reviewed studies evidence that the vast majority of heterogeneous FLPs are actually semi-heterogenized FLPs. As such, they consist of solids made around or functionalized with one FLP partner, while the second partner is non-covalently impregnated within the solid. Such semi-heterogenization strategies might lead to the leaching of one of the partners thus limiting the durability, and the subsequent productivity, of these however promising systems. The fully covalently heterogenized FLPs were based either on phosphine-functionalized metallated silica96 or on the physical mixture of two polymers functionalized respectively with one FLP partner,79 with however ill-defined and randomly distributed active sites as well as questionable utilization of the bulk FLP-functionalized solids. Furthermore, although the reported FLP-containing solids were demonstrated to efficiently activate either H2 or CO2, none of them were shown to perform direct CO2 hydrogenation so far.

3. Learnings from computational chemistry

The rational design and development of efficient FLP-based catalysts for CO2 hydrogenation requires a comprehensive understanding of the intricate relationships between molecular structures and their catalytic functionalities. This entails delving into the catalytic properties of molecular FLPs, while also rationalizing the conformational and environmental effects that govern their integration into the porous framework of materials. In this regard, quantum mechanical calculations have been extensively employed to: (i) obtain atomistic insights into the activation pathways of both H2 and CO2; (ii) identify the factors that influence their performances; and (iii) evaluate the potential impact of their immobilization inside porous matrices. This section covers the most relevant computationally-derived advances in this direction, ultimately aiming to predict and drive the synthesis of efficient FLP-based catalytic systems. Sections 3.1 and 3.2 summarize the main findings inferred from studies on molecular FLPs, while Section 3.3 delves into in silico studies on heterogenized FLP systems.

3.1. H2 splitting in molecular FLPs: a generally accepted initial step

The hydrogenation of CO2 by FLPs has been often shown to require the initial generation of an active hydrogenating species. This can be achieved via the heterolytic splitting of a H2 molecule into a hydride and a proton, which are accommodated on the LA and LB centers conforming the FLP, respectively. Beyond CO2 hydrogenation purposes, the splitting of H2 by FLPs represents an interesting process per se, playing a central role in the hydrogenation of a variety of unsaturated compounds.26 Besides, the ability to split H2 under mild conditions confers FLPs the potential to be used as molecular platforms to store H2 fuel in a safe and easy-to-handle manner. On these grounds, extensive theoretical have been devoted to understanding the reactivity of FLPs towards H2 and the factors that govern it. This section aims to give a brief overview of the most relevant computationally-derived findings in this sub-area, as well as the current controversies and challenges. For a more comprehensive review of computational studies on H2 splitting by FLPs, please refer to the ESI. Also, an extensive array of experimentally tested cases involving H2 splitting by FLPs has been explored employing computational methods (see ESI for details). These efforts have contributed to the establishment of a well-defined mechanistic picture depicted in Fig. 16 and are summarized as follows.
image file: d3cs00267e-f16.tif
Fig. 16 Overview of the main reaction mechanisms characterized by DFT calculations to govern the H2 splitting by (a) intramolecular and (b) intermolecular FLPs. 3D-structures on the right show representative transition-state geometries for: H2 activation by (c) intramolecular FLPs through an intramolecular path. Reprinted with permission from ref. 97. Copyright 2016, Wiley-VCH; (d) intramolecular FLPs through an intermolecular path. Adapted from ref. 98 with permission from the PCCP Owner Societies; (e) an intermolecular FLP. Reprinted with permission from ref. 99. Copyright 2008, Wiley-VCH.

On the one hand, intermolecular FLPs proceed via the formation of a non-covalent adduct between the LA and LB partners also referred as an “encounter complex”. This is followed by the insertion of H2 into the reactive pocket and its splitting to form a hydrogenated [LB–H]+⋯[LA–H] ion pair.

On the other hand, intramolecular FLPs may either activate H2 by themselves or dimerize to operate as an intermolecular FLP. The competition between intra- and intermolecular pathways in intramolecular FLPs has not been systematically investigated for a whole range of systems. Thus, setting clear conclusions about the factors that determine their relative likelihood remains challenging. Still, the prevalence of each of these paths might depend on the equilibrium LA⋯LB distance within the intramolecular FLP and on the ability of FLPs to dimerize into stable non-covalent adducts. Up to now, this mechanistic knowledge has been successfully applied to rationalize the H2 splitting activity of a wide variety of FLPs, including mostly P/B98–116 (known as thermally-induced FLPs53,117) N/B pairs,29,97,113,115,118–123 and less often carbene/B124 and N/TM (TM = Ni, Pt) pairs,125 or even that of heterogeneous systems such as hydroxylated indium oxide surfaces.126–128

Computational efforts have been devoted to identify the key factors that govern the H2 splitting, aiming at establishing structure–activity relationships and clear design rules. The primary structural requirement of an active FLP is to offer a suitable LA⋯LB distance. For instance, Vankova and co-workers determined that optimal P⋯B distances for H2 splitting lie within a range of 3 to 5 Å by performing constrained potential-energy surface scans along the P⋯B distance of a P(tBu)3/B(C6F5)3 FLP.129 As for B/N FLPs, Corminbœuf and co-workers identified that the optimal B⋯N distance for H2 splitting is ca. 2.9 Å. Still, the range of distances in which the energy of the TS varies within an energy range of 5 kcal mol−1 spans from ca. 2.6–3.5 Å.68 Thus, when targeting the heterogenization of FLP systems, it is essential to ensure that the positioning of the LA and LB partners within the material does not impose a suboptimal LA⋯LB separation.

Another key factor in FLPs’ reactivity is the electronic properties of the substituents of both LB and LA centers, affecting their basicity and acidity, respectively. A first attempt to set relationships between such molecular features of FLPs and the H2 splitting free energy (indicating its thermodynamic feasibility) was proposed by Pápai and co-workers.53 They carried out a comprehensive computational investigation using a series of intra- and intermolecular FLPs, which led to several conclusions, which were further supported by other works (see ESI for details):97,114,123,129

(i) The thermodynamics of the H2 splitting correlates with the cumulative acid–base strength of the FLP partners, which can be quantified from proton and hydride attachment energies;

(ii) Linked or intramolecular FLPs benefit from a smaller entropic penalty along the reaction coordinate than intermolecular FLPs and thus, require smaller cumulative acid–base strengths to be active;

(iii) Reaction free-energies for H2 splitting by the analyzed set of intramolecular FLPs was correlated inversely with the LA⋯LB distance. In other words, the shorter the donor–acceptor distance is, the more exergonic the H2 splitting is.

Ye and Johnson further proposed an original approach to explore computationally the immobilization of intramolecular FLPs for CO2 hydrogenation in UiO-66, a porous MOF matrix.89 They aimed to set the very first structure–activity relationships between the structure of molecular FLPs and their ability to activate H2 and hydrogenate CO2 in the H2/CO2 gas stream. Here, we focus on their findings on the initial H2 splitting step, while CO2 hydrogenation will be discussed in Section 3.3. The authors selected a set of 8 B/N FLPs, which on paper, can be covalently grafted on the terephthalate linkers of UiO-66 (Fig. 17a). These consist of a pending pyrazole functionalized with a BR2 moiety that leads to a series of B/N FLPs of increasing acidity by varying R. Computing electronic energy profiles for all systems, the authors tried to find Brønsted–Evans–Polanyi (BEP) relationships130,131 between energetic parameters (reaction energies and barriers) and a set of molecular descriptors. The latter include the acidity of the LA, the basicity of the LB, the electronegativity, chemical hardness and softness of the FLP sites, atomic charges and structural parameters of both the bare FLP and the zwitterionic hydrogenated intermediate.


image file: d3cs00267e-f17.tif
Fig. 17 (a) Octahedral cage of the UiO-66 metal–organic framework bearing a linker functionalized with a Lewis Pair (X). (b) Adsorption energy of H2 in UiO-66-X plotted against the free-energy of hydride attachment to the B center of the Lewis pair. Reprinted with permission from ref. 89. Copyright 2015, Wiley-VCH.

The H2 dissociative adsorption energy (i.e. heterolytic splitting of H2) was found to correlate linearly with the hydride attachment energy (Fig. 17b).89 As the LA strength increases, so does the H2 dissociative adsorption energy. Notably, no correlation was found with the proton attachment energy, in line with the experimental findings of Berke and co-workers.132 Still, we can note this computational work only addressed the impact of the substituents at boron (LA) center, resulting in a rather narrow range of proton attachment energies. The analyzed FLPs do not significantly differ in terms of LB basicity, leaving the impact of the latter uncharted.

Another linear BEP relationship was identified between the H2 splitting energy barrier and the LA chemical hardness.89 Increasing the hardness of the LA decreases the barrier for H2 splitting in accordance with Pearson's theory. We wish to stress here that, unlike the H2 adsorption energy, this barrier did not correlate with the acidity of the LA. The BEP principle thus does not fully apply to the H2 splitting process, given the absence of a linear correlation between reaction energies (thermodynamics) and barriers (kinetics). The other BEP relationship found between the H2 splitting barrier and the bond angle formed by Nb, Na and B centers in the bare FLP was ascribed to the fact that larger angles induce strain in the FLP, which lowers the barrier. It is also relevant for the present review that the MOF environment was found to marginally affect the adsorption energy of H2 on the FLP, as similar reaction energies were obtained from both isolated and in-MOF FLP models.

Two years later, Ye et al. revisited the factors that influence the thermodynamics and kinetics of H2 splitting on FLPs grafted in UiO-66 by enlarging their study to four families of intramolecular FLPs including B/N and B/P pairs attached to structurally distinct molecular scaffolds.91 Although quantitative linear correlations between H2 splitting energies and hydride attachment-free energies were found again for each FLP family (with r2 > 0.95), as shown above, they did not apply for all FLP families taken together (r2 = 0.834). This evidenced that besides hydride attachment-free energies, other factors influence the binding energy of H2. An in-depth screening of molecular descriptors revealed that it is also proportional to the LA⋯LB distance in the bare FLP and to the variation of the angles concerning the LA/LB sites and the two consecutive atoms of the scaffold in the direction of the FLP partner upon H2 binding.

Overall, despite the complexity of FLP's reactivity, computational methods have significantly contributed over the years to the generation of valuable knowledge that can be used to guide the experimental design of FLPs to conduct H2 splitting. Still, several points lack a quantitative unambiguous answer. For instance, the exact structural and electronic parameters that determine whether an intermolecular FLP operates through an intra- or an intermolecular mechanism are yet to be understood. Other aspects of the H2 splitting by FLPs that remain controversial or unclear nowadays are the relative weights of the impact of electric field polarization and orbital overlap on H2 splitting, as well as those of LA and LB strengths and how they relate to the FLP's nature and structure (see Fig. S6 and related text in ESI). Also, although a set of molecular parameters of FLPs have been recognized to influence H2 splitting kinetics and thermodynamics, generic and quantitative structure–activity relationships are still to be established. Most likely, this is because, besides eventual exceptions, the series of FLPs analyzed in each work are constrained to a single family of FLPs and quite often, account for a rather limited number of structures.

More importantly, within the present context of using FLPs not only for H2 splitting but also for subsequent hydrogenation of CO2, balancing the energetics of H2 splitting with those of the subsequent hydrogenation step is crucial. Specifically, FLPs should bind H2 strongly enough to make the H2 splitting step feasible, but not too strong so as to preclude subsequent hydride/proton transfer to an incoming CO2 molecule. In other words, a too strong Lewis base/acid would produce a too weak Brønsted acid/hydride donor, hampering CO2 hydrogenation. This adds an extra layer of complexity to the design challenge: the goal is drifted from minimizing energy barriers while favoring thermodynamics for each single elementary step process to identifying the molecular descriptors and the sometimes-overlooked experimental conditions that simultaneously optimize the performance of two distinct processes.

3.2. Molecular FLPs for CO2 hydrogenation

Carbon dioxide is an exceptionally stable molecule and its activation and conversion are hence not straightforward. The ability of FLPs to heterolytically split H2 and subsequently hydrogenate organic compounds has prompted the exploration of the potential use of FLPs for the direct hydrogenation of CO2 with H2 following the two-step process shown in Scheme 1.
image file: d3cs00267e-s1.tif
Scheme 1 FLP-mediated CO2 hydrogenation mechanism.

However, CO2 can be also captured by FLPs as shown in Section 1 and hence, compete with H2 for FLP sites.13 Notably, FLPs often bind CO2 more strongly than H2, which may hamper H2 splitting as the first step of CO2 hydrogenation because of an additional energy penalty associated to CO2 release. Still, such FLP–CO2 adducts are not necessarily detrimental for CO2 hydrogenation, as they could allow distinct mechanistic scenarios, as discussed in detail below.

3.2.1. Computational mechanistic studies on FLP-promoted CO2 hydrogenation. Early mechanistic computational studies were performed on systems that leverage the aforementioned ability of FLPs to capture CO2 to initiate reduction processes that yield CO133 or methoxy–boranes/methanol,62–65 using intermolecular Zn/P and intramolecular B/P FLPs, respectively. More related to the present review, Wen et al. used in 2013 DFT to investigate the hydrogenation of CO2 by [TMPH][HB(C6F5)3], the H2-activated product by the TMP/B(C6F5)3 FLP, (TMP = 2,2,6,6-tetramethyl-piperidine).134 The H+/H transfers from [TMPH][HB(C6F5)3] to CO2 were found to proceed in a concerted fashion through a single TS. A formic acid molecule is formed, which is subsequently deprotonated by TMP to yield [TMPH][HC(O)OB(C6F5)3], where the formate group is bound to the B center through an oxygen atom.

At this point, computations had already granted atomically-resolved mechanistic details of both elementary steps underlying CO2 hydrogenation (i.e. H2 splitting and H/H+ transfer). However, it was not until 2015 that Fontaine, Stephan and co-workers combined them both to study the overall hydrogenation mechanism of CO2 with H2 catalyzed by the intramolecular B/N FLPs, 1-BR2-2-NMe2-C6H4 (R = 2,4,6-Me3C6H2 and 2,4,5-Me3C6H2), as represented in Fig. 18a.32 These FLPs were experimentally found to afford the formation of boron-bound formates, acetals and methoxides, the product ratio being sensitive to the experimental conditions of H2/CO2 pressures. Traces of methane were also detected at low CO2 pressure.


image file: d3cs00267e-f18.tif
Fig. 18 Compilation of computationally studied FLP structures for CO2 hydrogenation with H2.

The authors then applied DFT calculations to investigate the initial hydrogenation steps.32 They include the hydrogenation of the bare FLP, followed by the formation of formic acid via concomitant H and H+ transfers to an incoming CO2 molecule. In line with previous calculations by Zimmerman et al.135 and Wen et al.134 the hydrogen transfer step was found to take place with a concerted TS. Thereby, the B-bound hydride is transferred to the electrophilic C atom of CO2, while one of the oxygen atoms of CO2 abstracts the acidic proton from the N center of the FLP. Interestingly, the analyzed FLPs (Fig. 18a) were observed to undergo protodeborylation after H2 splitting (Fig. 19a), giving access to 1-BHR-2-NMe2-C6H4 compounds. After a second H2 splitting step, the latter led to hydrogenated species analogous to that represented in Fig. 19b, which were found to hydrogenate CO2 through affordable free-energy barriers. Fig. 19b shows a representative TS structure for CO2 hydrogenation.32


image file: d3cs00267e-f19.tif
Fig. 19 (a) First protodeborylation process in hydrogenated forms of 1-BRH2-2-NHMe2-C6H4 (R = 2,4,6-Me3C6H2 and 2,4,5-Me3C6H2). (b) Optimized geometry of the TS for CO2 hydrogenation by 1-BH2(2,4,5- Me3C6H2)-2-NHMe2-C6H4, obtained at the ωB97X-D/6-31 + +G(d,p) level of theory, including benzene as solvent through the SMD continuum solvent model. Adapted from ref. 32. Copyright 2015, Royal Society of Chemistry.

In 2018, Jiang et al. studied computationally the hydrogenation of CO2 promoted by a series of B/P intramolecular FLPs represented in Fig. 18b–d.136 In addition to the previously established mechanism (H2 splitting followed by the concerted H/H+ transfer to CO2, see path (i), green arrows in Fig. 20), the authors explored a novel mechanism that starts with the activation of CO2 on the bare FLP (path (ii), purple arrows in Fig. 20). Then, metathesis of H2 with the LB–C bond leads to an intermediate bearing an LA–bound formate and a protonated LB center, which further evolves to release formic acid regenerating the FLP. In most cases, path (i) was claimed to be more favorable than path (ii), except for FLPs shown in Fig. 18b, structures (i) and (iv). Despite the possible contribution of the mechanism associated with path (ii) (Fig. 20), path (i) which involves sequential H2 splitting followed by hydride/proton transfer to CO2 has been consistently adopted in subsequent computational studies on FLP-catalyzed CO2 hydrogenation, including both inter-43,50,137 and intramolecular91,138,139 FLPs as well as classical Lewis pairs.42 There is an exception, though. For the intramolecular B/N FLP displayed in Fig. 18e, Ghara and Chattaraj explored alternative mechanisms whereby both H2 and CO2 are concomitantly activated by the FLP sites.138,140 However, these were found to involve rather high free-energy barriers, ranging from 37 to 50 kcal mol−1. Besides, since the formation of non-covalent FLP⋯H2 and FLP⋯CO2 adducts is generally endergonic, the trimolecular nature of the proposed TSs is expected to make them kinetically unlikely. Such features render these concerted H2 + CO2 activation mechanisms not able to compete with stepwise processes represented in Fig. 20, path (i), as later proved by the same authors.138


image file: d3cs00267e-f20.tif
Fig. 20 Proposed pathways for CO2 hydrogenation with intramolecular P/B FLPs.
3.2.2. Impact of FLP's structure on CO2 hydrogenation activity. Understanding how the chemical structure of FLPs influences their ability to hydrogenate CO2 is certainly more complex than rationalizing their H2 splitting performance. This is due to the intrinsic step-wise nature of the CO2 hydrogenation mechanism, as described in the previous section, which requires fine-tuning of the FLPs’ structure to simultaneously optimize its performance for two distinct processes, i.e. H2 splitting and H+/H transfer to CO2. This sub-section outlines the computational efforts that have been devoted so far to establishing relationships between FLPs’ features and their ability to promote CO2 hydrogenation. Note that most of the knowledge gained so far holds for FLPs that operate through path (i) (Fig. 20). Conversely, path (ii) (Fig. 20) has been only studied for a very limited number of FLPs and hence not enough data is available yet to attempt the development of design rules to promote this pathway.

In 2016, Liu et al. carried out a computational investigation aimed at finding relationships between the electronic structure of FLPs and the kinetics of CO2 hydrogenation into formic acid.50 In particular, their work begs important and timely questions. Is there any balance or relationship between the H2 splitting and the hydride/proton transfer to CO2 steps? Would a stronger FLP combination favor both steps or exclusively the H2 splitting step as detailed above, thus being detrimental to the kinetics of the hydride/proton transfer step? Seeking to answer these questions, the authors carried out DFT calculations to analyze the free-energy profiles for CO2 hydrogenation by five intermolecular B/N FLPs, based on combinations of B(C6F5)3 and B(p-C6F2F3)3 LAs with TMP, btam and Lut LBs (see Fig. 18f). In line with the conclusions inferred from Section 3.1, the kinetic analysis of the H2 splitting step indicates that the stronger the base or the acid, the lower the free-energy barrier. However, no systematic trend was found between the FLP strength and the barriers for hydride/proton transfer for CO2 hydrogenation. Still, stronger FLPs generally exhibit higher energy barriers for the hydride/proton transfer step assorted with later TS, thus following the opposite trend than that of the H2 splitting step. This was ascribed to the fact that stronger FLPs are more prone to “catch and hold” protons and hydrides, hampering in turn the subsequent hydrogenation of CO2. Later on, this phenomenon was further rationalized based on natural bond orbital and molecular electrostatic potential analyses as well as methods that are based on the application of electron localization functions.139 In short, these showed that lower occupancies of the p(B) orbital result in stronger donor–acceptor interactions upon H2 dissociation, thus easing H2 splitting but hampering CO2 hydrogenation. From a fundamental perspective, this can be interpreted as a strong (weak) Lewis base yields a weak (strong) Brønsted acid after H2 splitting. Similarly, a strong (weak) Lewis acid partner will provide a weak (strong) hydride donor. It is apparent that the factors favouring H2 splitting (strong LA and/or LB) might hinder the H+/H transfer to CO2.

The results described above paved the way for setting structure–activity relationships. Optimal FLPs for CO2 hydrogenation must balance their ability to split H2 with that of transferring a hydride and a proton from the zwitterionic LBH+/LAH intermediate to CO2, leading to a scenario where both steps take place through accessible free-energy barriers. Otherwise, too weak or too strong FLPs could suffer from too slow kinetics for H2 splitting (Fig. 21, purple arrow) or hydride/proton transfer to CO2 (Fig. 21, red arrow), respectively.


image file: d3cs00267e-f21.tif
Fig. 21 Schematic free-energy profile for the two-step hydrogenation of CO2 catalyzed by FLPs. The green profile illustrates an optimal situation where both H2 splitting are kinetically accessible. Conversely, grey-dashed profiles represent sub-optimal situations limited either by a slow H2 splitting step, for which the barrier is highlighted with a purple arrow; or by a slow hydride/proton transfer step, in which the associated barrier is highlighted with a red arrow.

Corminbœuf and co-workers recently made substantial achievements in this context. In 2022, they reported a comprehensive computational exploration of intermolecular B/N and B/P FLPs for CO2 hydrogenation into formate.43 A wide spectrum of chemical space was covered with a library of 60 N- and P-based LBs and 64 triaryl borane LAs partners (Fig. 22a). Upon mutual combinations, a total of 3840 FLP candidates were computationally screened for CO2 hydrogenation. The relative stabilities of the intermediates and TSs shown in the catalytic cycle of Fig. 22b were computed, considering the LA in catalytic concentrations and LB in excess to provide the driving force for product formation. From the above data set, two sub-pools were extracted fixing one of the Lewis partners and varying the complementary one. These sub-sets allowed exploring the individual impact of the properties of both Lewis partners. To this end, linear free-energy scaling relationships (LFESRs) were established to map the performance of the FLP on CO2 hydrogenation as a function of two chemical descriptors, i.e. the free energy of proton attachment (FEPA) to the LBs, and the free energy of hydride attachment (FEHA) to the LAs, which were estimated as the free energies of the following reactions (eqn (1) and (2), respectively):

 
LB + NH4+ → [LBH]+ + NH3; ΔGr = FEPA (1)
 
LA + H → [LAH]; ΔGr = FEHA (2)
FEHA and FEPA can be regarded as direct measures of the acid and base strength of the Lewis partners, successfully employed as descriptors in the field of FLP-promoted H2 splitting (see Section 3.1). DFT-derived turnover frequencies (TOFs) were calculated as the free energy difference between the turnover-determining intermediate (TDI) and the turnover-determining transition state (TDTS) and used as a response variable that accounts for the FLPs’ catalytic activity towards CO2 hydrogenation.


image file: d3cs00267e-f22.tif
Fig. 22 (a) Library of B-based LAs and N- and P-based LBs used to create a pool of intermolecular FLPs, whose potential for CO2 hydrogenation was computationally explored by Corminbœuf and co-workers. (b) Catalytic cycle considered for the generation of DFT-derived TOFs. Reprinted with permission from ref. 43. Copyright 2022, Wiley-VCH GmbH.

Analysis of the LFESRs over the two sub-sets defined in Fig. 22a yielded volcano plots relating the FEHA and FEPA descriptors to the calculated TOFs. These plots allowed the identification of different regions as a function of the nature of the rate-determining step as illustrated in Fig. 23b (i.e. the TDI/TDTS combination designating the TOF). On the one hand, using the rather weak B-Me4 LA partner, most of the screened FLPs face difficulties in splitting H2 through TS1 as defined in Fig. 22b. Strong amidine bases like DBU provide the optimal kinetic balance among the steps of the cycle, making this combination the most active. Still, the screening of the LAs using the weak pyridine LB as fixed partner revealed an over-stabilization of the boroformate complex 5 (as defined in Fig. 22b) when strong LAs with electron-withdrawing substituents are employed. Such over-stabilization is due to too strong B–O interactions. This adds energy penalty for splitting H2 reflected in the 5/TS1 identity of the TDS/TDTS combination, as the dissociation of 5 to regenerate the bare FLP is an uphill process.


image file: d3cs00267e-f23.tif
Fig. 23 (a) Map correlating the TOF (log scale) for CO2 hydrogenation into formate to FEPA and FEHA descriptors for the 1664 B/N intermolecular FLPs. (b) Map analogous to that in panel (a) in which each grid point is coloured according to the identity of the TDI and TDTS. (c) Chemical compositions of the FLP combinations (A)–(C). Their corresponding FEPA and FEHA descriptor values are shown below. (D) corresponds to a “mismatched combination” between the LB from (A) and LA from (B). (E) is a poorly active combination with strong LA and LB components. Reprinted with permission from ref. 43. Copyright 2022, Wiley-VCH.

Graphing both descriptors, FEHA and FEPA, against the calculated TOF revealed a dark-red colored region where the CO2 hydrogenation activity is maximized (Fig. 23a). Suboptimal performances in going from red to blue-colored regions arise either from over stabilization of intermediate 5 or from high-energy lying TSs, as shown in Fig. 23b. Importantly, this work revealed that the acidity and basicity of the components have to be appropriately balanced to attain efficient FLPs for CO2 hydrogenation, creating unprecedented mapping of FLPs chemical composition and linear scaling relationships.43

In the sub-field of intramolecular FLPs, the nature of the molecular scaffold or backbone that supports both LA and LB centers is an additional parameter that affects the energetics of CO2 hydrogenation. This does not only influence the electronic properties of FLP partners but also determines their interatomic distance, allowing for strategic control of the geometrical features of the FLP pocket, which have been demonstrated to govern its reactivity.68,129

Using a series of intramolecular “bridged” B/P FLPs (Fig. 18b–d), Jiang et al. investigated how the length and the chemical nature of the bridge between P and B centers influence their CO2 hydrogenation reactivity.136 For alkyl-bridged FLPs, decreasing the electron-donating nature of groups in the R′ position (near the B center) was found to significantly favor H2 splitting. Still, this goes along with the formation of too stable zwitterionic intermediates that hinder hydride/proton transfer kinetics for CO2 hydrogenation. Similarly, for intramolecular B/N FLPs shown in Fig. 18g, Ghara et al.138 reported that the free-energy barrier for H2 splitting decreases upon the incorporation of an amino group in the para position of the pyridine ring, while the barriers also decrease in the order X = NH > O > S. However, S-containing systems have more difficulties than the others in reducing CO2. Elongating the bridge length in intramolecular B/P FLPs (Fig. 18c) was found to have a detrimental impact on CO2 hydrogenation, as Jiang et al. reported a systematic increase of the barriers for both H2 splitting and CO2 hydrogenation steps due to an increase of the strain in the TSs.136 Vinyl bridges (Fig. 18d), were not identified as promising candidates neither because they yield very stable zwitterionic intermediates after H2 splitting, hindering hydride/proton transfer steps.136

The flexibility of the backbone, which impacts LA⋯LB separation, has been also identified as a relevant parameter to tune the reactivity of intramolecular B/P FLPs towards H2 and CO2 to produce formic acid.137 Inspired by the application of dimethylxanthene backbones to prevent the self-quenching of the FLP components,141 Delarmelina et al. analyzed in detail a series of xanthene-inspired scaffolds which display various P⋯B distances (Fig. 18h–l and Fig. 24).137 Compound H (Fig. 24) is found active for H2 splitting in line with experimental observations,141 but exhibits a very height barrier for the CO2 hydrogenation step. Compared to H, both compounds I and J (Fig. 24) show slightly higher but still affordable barriers for H2 splitting. However, the impact of their backbone on their CO2 hydrogenation step was found to be very different: Catalyst I, which displays a shorter P⋯B distance than J, has a very high barrier for the second step; conversely, in J, the more rigid backbone and longer P⋯B distance led to a substantial decrease of the free-energy barrier for hydride/proton transfer to CO2 when compared to H and I, which translates into a reduction in the activation energy for the whole process. Further tuning of the FLPs H–J via modification of the electronic properties of the substituents at the B and P groups was found to play a secondary role compared to the stiffness of the backbone. Of note, Delarmelina et al. also reported that more polar solvents stabilize both the zwitterionic intermediate and the TS of the second step. Still, such stabilization is very similar in both cases, having an overall negligible impact on the height of the CO2 hydrogenation barrier.137


image file: d3cs00267e-f24.tif
Fig. 24 P⋯B interatomic distances (Å) between P- and B-centers of FLPs H, I and J studied computationally by Delarmelina et al.137 Adapted with permission from ref. 137. Copyright 2021, Elsevier.
3.2.3. Computational analysis of potential poisoning and inhibition routes. As described in the previous section, CO2 hydrogenation is commonly assumed to require a perfectly orchestrated sequence of H2 splitting followed by hydride/proton transfer to a CO2 molecule to yield formic acid, regenerating the bare FLP and thus, closing the catalytic cycle. However, the topology of potential energy surfaces can sometimes favor side processes that may compromise the efficiency of the targeted process. Besides comprehensively analyzing CO2 hydrogenation pathways and the factors that govern them, computational techniques have been also employed to identify and explore potential poisoning routes or side processes that can be detrimental to CO2 hydrogenation.

Owing to the well-known ability of FLPs to activate CO2-forming LB–C(O)O–LA adducts,54,142–144 the competition between H2 and CO2 for the binding sites of FLPs is intrinsic to the conditions required for catalytic CO2 hydrogenation, in which the FLPs are exposed to a mixture of both reagents. Thus, depending on the thermodynamics and kinetics of CO2 binding, the latter may be regarded as a potential poisoning route, precluding CO2-hydrogenation reactivity. If the formation of a covalent FLP–CO2 complex is exergonic and faster than H2 splitting (Fig. 25, light-blue dashed lines), the H2 splitting process will be slowed down. This is because the overall barrier that needs to be overcome corresponds to the free-energy difference between the FLP–CO2 complex and the TS1 for H2 splitting, as exemplified in Fig. 25Goverall,1, light blue arrow). In the worst-case scenario, this can even inhibit H2 splitting (and subsequent CO2 hydrogenation) if CO2 binding is irreversible or if the resulting overall barrier is too high to be overcome at the experimental conditions. Still, it should be noted that for systems that can operate through path (ii) (Fig. 20) thanks to a low FLP–CO2 → TS1′ barrier, the preferred binding of CO2 with respect to H2 splitting might not lead to a poisoning route but rather to an in-cycle species. However, the factors that determine the preference for one path or another remain elusive, making it difficult to predict whether a given FLP would be able to bypass the limitations underlying competitive CO2 binding processes by proceeding through this alternative mechanism.


image file: d3cs00267e-f25.tif
Fig. 25 Schematic free energy profile for the early steps of CO2 hydrogenation, highlighting the impact of possible stabilizing interactions between the bare FLP and CO2 (light-blue dashed lines) or formic acid (red dashed lines) on the kinetics of H2 splitting. Black, light-blue and red arrows represent the individual free-energy barrier for the H2 splitting step (ΔG), and overall barriers for the same process measured from stable FLP–CO2Goverall,1) and FLP–HCOOH (ΔGoverall,2) complexes, respectively. TS1’ accounts for possible hydrogenation of CO2 through path (ii) of Fig. 20.

Assuming path (i) in Fig. 20 (sequential H2 splitting followed by H+/H transfer to CO2) as the mechanism at work, computational studies aimed to identify the factors that determine the relative affinity of FLPs toward H2 and CO2,91,121,145 hence providing guidance to favor H2 over CO2 binding. Changing the electronic properties of FLP sites via tuning of LA/LB substituents of structures shown in Fig. 18m–q was only found to either strengthen or weaken the adsorption of both H2 and CO2. Conversely, shortening the LA⋯LB distance can simultaneously strengthen the H2 adsorption and weaken the CO2 adsorption.91 Also, thermodynamics of H2 and CO2 binding can be controlled by playing with the electron-conjugation properties of FLPs. In particular, gains in aromaticity of the FLP structure along CO2 binding were found to result in lower activation energies and more enhanced exergonicity due to the stabilization of TS and product structures.121 On the other hand, reduction of anti-aromaticity along H2 splitting is also beneficial in terms of kinetics and thermodynamics.145 Thus, although not straightforward to control, these properties may be also tapped to achieve tailored H2/CO2 binding selectivity.

Akin CO2 or H2, the formic acid product resulting from CO2 hydrogenation can also form strong covalent adducts with bare FLPs, as illustrated in Fig. 25 (red dashed lines), which could hamper the subsequent H2 splitting process to initiate a new cycle. One of the rare examples of computational works that considered the participation of FLP–HCOOH complexes as in-cycle species for CO2 hydrogenation (Fig. 22b) revealed that these complexes act as the resting state of the catalyst in a wide range of LA/LB combinations (Fig. 23).43 Comparable occurrences may be also expected for further reduced species such as methanol. The latter has been reported to strongly bind boron LA centers of FLPs forming B-methoxy species, both computationally134 and experimentally.62–65 To circumvent these issues, as well as to provide the thermodynamic driving force for the per se endergonic CO2 hydrogenation reaction, bulky bases such as DBU may be used to “trap” the HCOOH product as [Hbase][HCOO] salts, while leaving unaltered LA sites of FLPs by avoiding quenching through steric repulsion (see Table 1).

Finally, as the vast majority of FLPs rely of boron LA centers, common FLPs are highly sensitive to moisture due to the oxophilicity of boron atoms. Moreover, they can also suffer from other degradation or inhibition pathways, such as protodeborylation, as shown in Section 3.2.1 (Fig. 19a).32 conspicuously, this could either promote the complete degradation of the FLP or lead to partially protodeborylated species with enhanced activity. DFT calculations suggested that the ability of FLPs to undergo protodeborylation may be modulated by tuning the steric hinderance around the B–C bonds, so that crowded environments hamper this process, allowing control over the electronic and steric properties around boron LA sites during the course of the reaction. In general, this process is commonly overlooked in the literature, although its assessment appears to be necessary to evaluate or predict the activity of an FLP towards CO2 hydrogenation. Otherwise, FLPs that may be initially predicted to be inactive could become active upon protodeborylation and vice versa.

3.2.4. Concluding remarks. This entire section demonstrates that achieving efficient FLP catalysts for CO2 hydrogenation requires a delicate balance between the FLP's capability to cleave H2 and the ability of the resulting hydrogenated FLP to transfer a proton and a hydride to CO2 while also preventing or minimizing potential routes of poisoning. This requires granting an appropriate CO2/H2 affinity towards FLP sites and allowing the FLP–HCOOH complex to liberate the product.

Most of the knowledge gained from the computational study of molecular FLPs may apply to the design of heterogenized FLPs inside porous matrices for the sake of avoiding the limitations inherent to homogeneous systems. However, such a heterogenization strategy poses a series of additional challenges, related to the stability of the functionalized materials, the dynamics and accessibility of FLPs within the pores of the material, the selectivity of the material towards CO2/H2 adsorption, or the impact of confined environments on the electronic and structural properties of FLPs, as discussed in the following section.

3.3. FLPs heterogenized into porous solids for CO2 hydrogenation

3.3.1. Immobilizing Intramolecular FLPs in MOFs. So far, only a handful number of computational studies have considered the full immobilization of intramolecular FLPs in MOFs for CO2 hydrogenation. In these reports, the role of the pristine MOF consists of an inert support, i.e. with no participation of either organic linkers or inorganic subunits in the reaction. The catalytic activity of the molecular FLP is thus “transferred” to the porous solid through its covalent grafting to one of the MOF's organic linkers, allowing the in silico evaluation of its stability and reactivity in the resulting FLP@MOF solid. So far, such studies have been mostly devoted to the zirconium-based UiOs (University of Oslo) subfamily of MOFs.146 Such focus on UiOs is directly related to their experimentally well-known chemical and thermal robustness and the possible post-synthetic functionalization of their organic linkers.147

A couple of computational studies have considered the possibility that Zr-based MOF may possess built-in frustration between LA and LB centers. Along that line, Slater et al. have conducted extensive ab initio molecular dynamics (AIMD) simulations on the defective Zr-based UiO-66.148 They show that missing terephthalate linkers are charge-balanced by hydroxide anions bonded to under-coordinated Zr sites, where rapid proton shuttling may be involved with physisorbed atmospheric water molecules. At high activation temperatures, AIMD simulations further reveal that FLP sites may form consisting of an undercoordinated Zr site (Lewis acid) adjacent to a hydroxide bonded to a proximal undercoordinated Zr atom (Lewis base). The authors suggest that similar defects may exist in a wide range of MOFs, whereby increased catalytic activity and tailoring of the functional behavior could be targeted.

More recently, a detailed DFT study considered the mechanistic details of CO2 hydrogenation to methanol in the linker-defective UiO-66, exploring various pathways and free-energy profiles on the resulting metal-based FLPs.149 The defective model of UiO-66 is created by removing an organic carboxylate linker from a Zr6-oxocluster node, leaving behind two adjacent undercoordinated Zr sites. One of them is proposed to act as a Lewis acid, while the –OH group added to the other Zr site for charge balance purposes acts as a Lewis base. The hydrogenation consists of a three-stage transformation: (i) CO2 is hydrogenated into formic acid (HCOOH); (ii) HCOOH is converted to formaldehyde (HCHO) via hydrogenation and dehydration; (iii) HCHO is hydrogenated into CH3OH. For CO2 hydrogenation to HCOOH, three possible pathways were investigated (Fig. 26). Typically, in pathway III, H2 is initially located above the open Zr site (LA) and the hydroxide (LB) and dissociates, leading to a Zr–H hydridic site and a vicinal Zr–H2O water molecule. CO2 is then hydrogenated into formic acid (FA) via a slightly endergonic concerted mechanism overcoming an energy barrier of 10.2 kcal mol−1. The cis-FA desorbs to the gas phase and then transforms into the trans-FA. HCOOH is then converted at the FLP in a stepwise fashion into formaldehyde HCHO and CH3OH via facile H2 dissociation and the concerted H+/H transfer to HCOOH and HCHO. Interestingly, the authors explored the impact of a possible confinement effect in UiO-66 on the catalytic activity by considering two sizes (small/large) of clusters. It was found the energy barriers obtained with the two types of clusters differ only by 2 kcal mol−1, suggesting that the confinement effect in UiO-66 is negligible.


image file: d3cs00267e-f26.tif
Fig. 26 Proposed pathways for CO2 hydrogenation to HCOOH in the linker-defective UiO-66. Reprinted with permission from ref. 149 from the Royal Society of Chemistry.

Leaving apart these specific cases where the pristine MOF was studied for its intrinsic “FLP-like” catalytic properties for CO2 reduction,149 we are not aware of computational studies of functionalized FLP@MOF systems where the MOF itself plays a catalytic role as a Lewis partner for CO2 conversion.

Johnson and Ye pioneered the design of FLP@MOF by covalently grafting an intramolecular FLP to the terephthalate linker of UiO-6688,89 or embedding the FLP in UiO-6790 to study catalytic CO2 hydrogenation mechanisms. They aimed to functionalize UiO-66 with 1-(difluoroboranyl)-4-methyl-1H-pyrazole (P-BF2), developed to mimic 1-[bis(pentafluorophenyl)boryl]-3,5-ditert-butyl-1H-pyrazole FLP, which is known to cleave H2 heterolytically and fixe CO2.150,151 The less bulky P-BF2 was modelled by removing the tert-butyl groups and replacing C6F5 moieties by F atoms. Constructing the UiO-66-P-BF2 in silico with one P-BF2 group per primitive cell, static-DFT calculations in the gas phase provided relative energy profiles for CO2 chemisorption, H2 dissociation, and CO2 hydrogenation to formic acid, without accounting for entropic contributions.88 Two pathways were compared: CO2 activation followed by H2 cleavage or concerted H2 cleavage and CO2 hydrogenation. The first pathway involving the chemisorption of CO2 in UiO-66-P-BF2 lead to the formation of a very stable adduct and a potential poisoning of the FLP. The second pathway was found energetically more favourable than the first with a much lower activation barrier, making it the desired one. The MOF environment slightly reduced the barriers for H2 dissociation and CO2 hydrogenation compared to the isolated P-BF2 catalyst, though the absence of entropic contributions suggests caution.

Expanding on their initial proof-of-concept, the authors explored UiO-66-P-BR2 variants for CO2 hydrogenation,89 testing various electron-donating and -withdrawing substituents on the acidic partner while keeping the same diazole basic partner (see Fig. 17a). The process in the gas phase involved H2 heterolytic splitting (as presented in Section 3.1), CO2 chemisorption, and concerted hydride and proton transfer to form formic acid. They again identified BEP relationships for rapid screening of functional acid groups, finding that energy barriers for CO2 hydrogenation correlated linearly with H2 dissociation energy in the UiO-66-P-BR2 series (Fig. 27a). Weaker H atom bonds consistently facilitated H/H+ transfers, easing CO2 hydrogenation. H2 splitting energies served as a first estimate for CO2 hydrogenation barriers. A Sabatier activity map combined BEP relationships for H2 dissociation barriers (i.e. B's hardness) and CO2 hydrogenation barriers (i.e. H2 dissociation energy), revealing optimal functional groups with hardness above 3.8 and H2 binding energies between −0.45 and −0.1 eV (red area in Fig. 27b).


image file: d3cs00267e-f27.tif
Fig. 27 (a) BEP relationship between CO2 hydrogenation barriers and dissociation energies of H2 in UiO-66-P-BR2. (b) Sabatier activity map for overall CO2 hydrogenation at 298 K. The CO2 hydrogenation barriers are given by a BEP relationship with H2 dissociation (adsorption) energies (x-axis) and H2 dissociation barriers have a BEP scaling with the hardness of the LPs. Reprinted with permission from ref. 89. Copyright 2015, American Chemical Society.

So far, we are not aware of any experimental report on the functionalization of UiO-66 with P-BF2. The functionalization of this well-known MOF with FLPs is yet to be reported and may be hindered by practical synthetic limitations, discussed above in Section 2.2. Notably, Johnson et al. were the first to have screened the reactivity of FLPs anchored in MOF and attempted to find descriptors for their catalytic activity in CO2 hydrogenation. Still, descriptors related to the MOF's impact as a “secondary coordination sphere” of the FLP on the CO2 hydrogenation-free energy profile are yet to be identified.

Along this line, the same authors have compared the impact of pore size and topologies of MOFs on the dissociation of H2, the chemisorption of CO2 and its reduction into formate, examining MOFs with increasing pore sizes: MIL-140B (∼8 Å) < UiO-66 (∼5 Å and ∼9 Å) < MIL-140C (∼10 Å) < UiO-67 (∼12 Å and ∼16 Å).92 The Lewis pairs are comprised of N and BF2 moieties linked via a C atom, embedded in the organic linker, except UiO-67 where the LP is part of a side chain pointing towards the pore center, generating minimal steric interaction with the hosting framework. The calculated dissociation energies of H2 are found similar in all systems, suggesting that pore size and confinement effects have no significant impact because of the small size of H2. In contrast, CO2 chemisorption energies exhibit significant differences among the four MOFs. They depend on the orientation of the bound CO2 and the proximity of carboxylate O atoms of nearby linkers, which may result in strong repulsions with CO2. The strongest chemisorption of CO2 observed in UiO-66-P-BF2 is allowed by the minimal steric interactions between the LP and this MOF. By contrast, the lower strengths of CO2 chemisorption in the three other systems are a subtle convolution of the flexibility of the MOF's linker and specific confinement effects around CO2 imposed by the number of linkers per secondary building units. Turning to the CO2 hydrogenation process, the authors claim that confinement and steric hindrance may allow the selection of specific MOF topologies to (i) favor the dissociation of H2 over the chemisorption of CO2, thus increasing the resistance of the MOF to CO2 poisoning, and (ii) improve the catalytic performances by pre-activating CO2 via a bent physisorbed configuration.

Johnson et al. also considered UiO-67 (Fig. 28a) – due to its larger pore volume and surface area than that of the isostructural UiO-66 – for hosting an intramolecular FLP, namely UiO-67-(NBF2)4 (Fig. 28b).90 Notably, the adsorption energy of CO2 was found to be much lower than that of H2 (−0.22 eV versus −0.50 eV). This allows a much easier desorption of chemisorbed CO2 and avoids the poisoning of the FLP by CO2 that was previously observed in UiO-66-P-BF2. Overall, UiO-67-(NBF2)4 provides multiple LP catalytic sites for H2 dissociation, making facile reduction of CO2 to methanol possible without requiring diffusion of the intermediate reactants or products into adjacent pores. The reaction pathways to methanol involve a series of concerted two-hydrogen transfer steps where the following pathway, CO2cis-HCOOH → CH2(OH)2 → CH2O → CH3OH, has the lowest potential energy surface, HCOOH acting as a proton shuttle.


image file: d3cs00267e-f28.tif
Fig. 28 (a) Representation of the native UiO-67 and its biphenyl linker and (b) the modified UiO-67 incorporating a ligand with a B/N intramolecular FLP.

Heshmat reinvestigated in depth the above UiO-66-P-BF2 catalyst using ab initio molecular dynamics (AIMD) simulations, computing free energy surfaces (FESs) using metadynamics.93 Here, the thermal and entropic contributions to the free energy profile were taken into account thanks to the motion of H2/CO2 molecules in the gas phase within the LP-functionalized MOF. The mechanism of CO2 hydrogenation towards formate was shown to be more eventful than initially proposed by Johnson et al.,89 with the possibility of a stepwise mechanism in addition to the concerted one, while providing unique features about the dynamics of the reaction.

The free-energy barrier for H2 splitting was found to be affordable (18.8 kcal mol−1), being ca. 6 kcal mol−1 higher than the reported electronic energy barrier reported in Johnson's work,89 due to the incorporation of entropic effects. Also, simulations revealed that the formation of the zwitterionic intermediate (hydrogenated FLP) is thermodynamically favorable by ca. 6 kcal mol−1. Importantly, although the free-energy barrier for CO2 activation by the bare FLP was estimated to be 10 kcal mol−1, being lower than that for H2 splitting, the reverse free-energy barrier for CO2 desorption was computed to be affordable as well, accounting for 21.1 kcal mol−1.93 This suggests that even though CO2 binds stronger to the FLP than H2, the formation of FLP–CO2 adducts does not necessarily imply poisoning of the catalyst. Still, in the presence of a CO2/H2 mixture, H2 splitting would be precluded by an overall free-energy barrier close to 30 kcal mol−1 (H2 splitting barrier + energy span associated to CO2 desorption from the FLP). On these grounds, it is postulated that suitable CO2 hydrogenation conditions might involve exposing the MOF sequentially to separate streams of H2 (to activate it on the FLP) and CO2 (to hydrogenate it), rather than to a mixture of CO2 and H2, whereby the binding of CO2 to FLP sites would hamper H2 splitting.

The FESs for the next CO2 hydrogenation step were calculated using cis and trans conformations of dissociated H2. The cis conformation reveals an asynchronous concerted mechanism, while the trans conformation shows a stepwise mechanism with H is transferred to C(CO2) first, followed by H+ to O(CO2) after 400 fs. Both mechanisms are kinetically feasible and endergonic. The hydrogenation of CO2 may be thus fully described using the two H⋯B and H⋯C(CO2) distances as collective variables. Atomic charges analysis indicates a more polarized B–H bond vs. the N–H one, causing an earlier hydride than proton donation. This is also reflected in the 1 eV higher HOMO–LUMO energy gap between the lone pair of O(CO2) and N–H σ* than that between the B–H σ orbital and π* of activated CO2. MD simulations reveal strong interactions between the HCOOH product and Zr-clusters of the MOF, in line with known observations in Zr-based MOFs,152 implying the need for continuous HCOOH removal for high conversion rates.

Another recent computational study investigated the Zr-based UiO-66 MOF functionalized with alanine boronic acid for CO2 hydrogenation.153 The energy profile for CO2 hydrogenation in the gas phase of the immobilized FLP in UiO-66 was compared to that of the isolated molecular FLP. These calculations suggested an overall stabilization along the whole energy profile as the reaction proceeds in the FLP@UiO-66 solid, although entropic contributions were not considered. A notable decrease of 4.6 kcal mol−1 in the H2 splitting barrier was estimated upon immobilization of the FLP. Also, the comparison of the various components of the energy decomposition analysis (EDA)154 (electrostatic, Pauli repulsion, dispersion and orbital interactions) revealed a greater stabilization of reaction intermediates and transition states in the FLP@MOF system when compared to that obtained with the isolated FLP, pointing towards a beneficial impact of the confinement of the reaction within the MOF's pores.

3.3.2. Impact of immobilizing intermolecular FLPs on catalytic activities. The previous section focused on the targeted immobilization of a single intramolecular FLP into a porous inert host. Immobilizing both LA and LB partners as independent molecular species within the pores or channels of a MOF poses a more challenging target. This requires functionalizing the host with two chemically distinct species instead of just one. Still, it might be considered as a more attractive approach than the immobilization of a single intramolecular FLP for several reasons. Firstly, the successful anchorage of both the Lewis acids (LA) and Lewis bases (LB) to the MOF could be used to prevent the quenching of their reactivity. Secondly, the separate anchorage of the LA and LB partners at distinct positions of the MOF opens – in theory – the possibility of controlling the LA⋯LB distance, which may not be attainable in homogeneous conditions. Lastly, such strategy significantly increases the number of possible LA/LB combinations to explore when compared to those accessible in homogeneous conditions, thanks to the prevention of quenching.

However, there are key issues that need to be addressed in order to fully benefit from the aforementioned advantages of the immobilization of intermolecular FLPs. Firstly, it is crucial to control the synthetic or post-synthetic modifications of the MOF to achieve the desired relative positioning of the LA and LB centers within the required distance. Secondly, it is required to preserve the steric accessibility and chemical environment of both the LA and LB partners once anchored, in order to avoid any detrimental interactions between the LA or LB partners and the MOF. This raises yet-to-be-answered questions regarding the dynamics of both the MOF and the LA/LB partners. Lastly, it requires identifying the most suitable MOF structures from the vast number of existing ones that can meet the aforementioned constraints, while also considering the practical issues related to the compatibility of their synthesis protocols, thermal and chemical stability with those of the FLPs to be immobilized. In that respect, computational chemistry approaches have a key role to play, not only for exploring the potential catalytic performances of LA⋯LB pairs once immobilized within the MOF, but also for predicting the overall impact of their immobilization in the MOF pores or channels. The latter includes considerations on LA and LB's spatial partitioning, steric, chemical or electronic features and dynamical behavior within the MOF, to ultimately identify design principles.

In that direction, Corminbœuf et al. explored the constrained spatial arrangements of the prototypical BF3/pyridine Lewis pair.68 Using the B⋯N distance (d) and the angle (Φ) formed by the B's empty orbital and the N's lone pair as descriptors, the complete free energy profile for CO2 hydrogenation for intermediates and transition states as defined in Fig. 22b was evaluated at each increment of the B⋯N distance (0.1 Å steps), while allowing Φ to adapt freely (Fig. 29a). Fitting it to a Morse potential, a scaling relationship was created that predicts the reaction's rate or TOF (turn over frequency) based solely on the B⋯N distance. This allowed them to estimate the resulting gain or loss of catalytic performance of the « immobilized » FLP to the unconstrained one (ΔTOF), thus providing direct insight into how restricting the FLP's geometry affects its activity. The complete activity profile was constructed as depicted in Fig. 29b.


image file: d3cs00267e-f29.tif
Fig. 29 FLP-catalyzed hydrogenation of CO2 following the catalytic cycle described in Fig. 22b. (a) Energy–distance scaling relationship showing the variation of free-energies of each intermediate and transition state with the d descriptor (d = LA⋯LB separation in Å), fitted with Morse potentials (dashed lines); (b) Activity map describing the relative TOF of constrained FLPs with respect to unconstrained FLPs as a function of the key descriptors, d and Φ; (c) Intramolecular FLPs that are mapped in (b) and the geometric descriptor values extracted from the corresponding intermediate 2. Copyright 2022 from ref. 68. Angewandte Chemie International Edition published by Wiley-VCH GmbH under the terms of the Creative Commons license.

Remarkably, these calculations reveal that the catalytic activity of an N/B FLP may be boosted by constraining its geometry. The peak performance (dark red area in Fig. 29b) is estimated to be eleven orders of magnitude higher than that of the unconstrained FLP (ΔTOF = 0). High efficiency can be achieved over a wide range of distances as long as the N⋯B orientation is suitable. The optimal orientation for H2 activation is in the range of Φ = 90°–135°, which suggests that promoting this step through immobilization could enhance the catalyst's efficiency. Notably, the predictive capability of the activity map was successfully validated against reported intramolecular N/B FLPs and known experimental trends from Fontaine et al. on amine-boranes32 for CO2 hydrogenation.

This pioneering work of Courminbœuf et al. presents a comprehensive and generic conceptual approach to immobilizing FLPs in any type of solid-state system, regardless of structure and chemical environment provided by the host, as the explored geometric constraints closely resemble those that arise from FLP's immobilization. Although the specific influence of the host on the FLPs’ chemistry is not addressed, these findings convincingly demonstrate that constrained geometry correlates with a distinct catalytic activity. The strategic confinement of the FLP within specific distances and orientations can unexpectedly trigger enhanced catalytic activity when compared to the unconstrained FLP. Overall, this proposed approach has the potential to inspire future investigations for screening the catalytic activities of other chemical LA–LB partners in CO2 hydrogenation and identifying optimal constrained geometries to be further incorporated into MOFs or porous polymers for instance.

4. Lessons learned from experimental and computational studies

With tremendous achievements in both the experimental discovery of main group elements FLP systems and the computational understanding of their reactivity, the careful analysis of these two prolific sides of such a subtle chemistry opens opportunities for installing a dialogue between experiments and calculations, exemplified for the CO2 hydrogenation reaction.

4.1. Finding the sweet spot of reactivity with H2 and CO2

H2 splitting and CO2 hydrogenation processes generally follow opposite trends with variations of most of the FLPs’ properties, whereby whatever favors one is detrimental to the other. It is thus inevitable to outline parallelisms with the Sabatier principle, widely used in heterogeneous catalysis to describe an optimal catalyst that binds substrate molecules strong enough to enable reactivity but not too much to prevent undesired poisoning issues. Hence, pinpointing optimal FLP catalysts for CO2 hydrogenation with H2 demands the identification of sweet spots that balance the feasibility of both H2 splitting and subsequent CO2 hydrogenation step.

In this regard, the development of computationally driven strategies to identify sweet spots where the energetics of both steps are balanced to achieve optimal CO2 hydrogenation activity represents a matter of current research, and some guidance and rules of thumb have been recently reported. Specifically, optimal CO2 hydrogenation activity has been found to require a moderate cumulative acid/base strength, which can be estimated as the sum of proton and hydride attachment energies to the LA/LB partners (FEPA and FEHA, respectively). Thus far, systematic explorations have been only carried out with intermolecular B/N FLPs, which revealed that the TOF is maximized for LA/LB combinations with FEPA/FEHA ratios ranging linearly from ca. −35/−10 to +25/−35 kcal mol−1.43 Moving away from these ranges leads to too strong pairs, which form too stable zwitterionic hydrogenated FLP intermediates or bind products too strongly, or to too weak pairs, which have difficulties splitting H2.

Also, intramolecular FLPs allow for some “tricks” to move orthogonally away from these trends, by controlling either the LA⋯LB distances or the rigidity of the molecular scaffolds, as exemplified by the work by Delarmelina et al.137 More rigid backbones and longer LA⋯LB distances (of ∼4.4 Å) were found to substantially reduce the free-energy barrier for hydride/proton transfer to CO2, compared to more flexible scaffolds with LA⋯LB distances of <4 Å. While for H2 splitting, optimal B⋯N distances were found to lie in a range of 2.6–3.5 Å,68 not a single study has claimed yet what is exactly the optimal LA⋯LB distance to promote CO2 hydrogenation and how it varies as function of the nature of the FLP sites. Still, it is reasonable to assume that optimal performances would be found at the overlap of distance ranges that are suitable for both H2 splitting and H+/H transfer elementary steps.

In addition, potential poisoning pathways such as competitive binding of species on FLP sites or protodeborylation, are analyzed independently in the literature, but rarely (if ever) in a joint manner to grant a full mechanistic picture of the reaction. Regarding the latter routes, some trends have been found for specific sets of FLPs. These include: shortening of LA⋯LB distances to favor the activation of H2 by FLPs over that of CO2; or increasing the steric hinderance around B sites to prevent undesired protodeborylation events.

Moreover, beyond the determination of the adequate energetics balance and the inclusion of poisoning in computational approaches, the calculated data should be translated into experimental parameters such as partial gas pressure and temperature, which remain crucial towards applicable CO2 hydrogenation systems. For instance, a pioneering system by Corminbœuf and Dyson allows for the FLP-driven CO2 reduction at 120 °C under 130 bars, specific conditions that were not anticipated from the calculated data but must have required intensive experimental screening.43

4.2. Challenges in turning computational chemistry into target synthesis

Recently, computational chemistry approaches have become an extremely powerful tool for screening the potential use of molecular FLPs for CO2 hydrogenation, identifying successful LA–LB pairs, which were later verified experimentally. Still, the translation of computational learning into the experimental realm remains far from trivial, especially when tackling the challenge of FLPs’ heterogenization.

Two experimental aspects have to be considered when turning computational results into bench-scale trials: the synthetic accessibility of the FLP synthons and the conditions to be applied to reach the targeted FLP's reactivity. Considering the reactivity of computationally designed FLPs for the CO2 hydrogenation, the conditions reported in the two recent examples by Zhao42 and Corminbœuf,43 both claiming for a catalytic FLP system, were 160 °C under 60 bars of CO2/H2 and 120 °C under 130 bars, respectively. Far from being trivial, the findings of such harsh conditions to assess the efficiency of carefully designed Lewis pairs can only rely on a large screening of pressures and temperatures, including a range of values not often investigated for molecular catalysis. Thus, the quest for novel experimental FLP systems for CO2 hydrogenation, whether predicted by preliminary computational studies or not, must encompass from its beginnings these specific ranges of temperatures and pressures, especially when compared to transition-metal-based counterpart systems operating under smoother conditions.155

Besides the deployment of dedicated organic synthesis of the predicted FLPs, limitations that are related to their heterogenization may emanate from the reactivity/stability of the porous solids (MOF, POP, COF) themselves upon their functionalization with the targeted FLPs. Additional drastic limitations may arise from the catalytic conditions themselves (pressure, temperature, solvent…). In the proposed two-step synthetic pathway devised by Ye and Johnson90 for targeting the MOF-heterogenized FLP UiO-67-P-BF2, the initial step hinges upon a palladium-catalyzed Negishi cross-coupling reaction involving the commercially available 3-iodo-1H-pyrazole and a benzylzinc bromide derivative, which has been reported to yield up to 82%.156 However, it is worth noting that the synthesis of the 2,4,5-trimethylphenylzinc bromide coupling partner, though commercially procurable, is protected by patents and thus not readily accessible for research groups. Moving on to the second potential step, it entails the oxidation of p-xylene to terephthalic acid following the Amoco process.157 It is noteworthy that this oxidation process employs acetic acid as a solvent, compressed oxygen at approximately 200 °C, and a combination of cobalt, manganese, and bromide catalysts. These rigorous synthetic conditions raise concerns regarding the potential undesired oxidation of the benzylic position within the linker precursor, consequently leading to the decomposition of the intended motif.158 Lastly, the borylation of the pyrazole moiety has been reported utilizing a two-step methodology, which involves the quantitative utilization of an additional BCF:carbene adduct.150 Furthermore, it is important to acknowledge that this step relies on heterogeneous solid–liquid reactivity occurring at the interface of the MOF. It also requires to simultaneously ensure the stability of the HBF2 reactant against the reactivity of the Zr–OH groups present on the MOF's surface, as well as any pending carboxylic acid defects. This presents a significant challenge due to the potential risk of hydrolysis159,160 and the subsequent release of highly toxic hydrofluoric acid and fluoroboric acid, particularly given the potential diffusion limitations within the pores of the MOF.

Anticipating these synthetic limitations by computational approaches is extremely challenging. Consequently, the discovery of heterogenized FLP systems currently relies on the “trial-and-error” explorations of porous matrices from a broad range of potential candidates. However, the knowledge gained from both homogenous systems and theoretically derived structure–activity relationships can be used to narrow the range of FLP candidates to be immobilized, thus saving time consuming experimental efforts. Future work may explore potential synergies between the chemistry of organic linkers and their functionalization, and that of metal nodes in MOFs72,73 towards the construction of heterogeneous FLPs, which are still rather scarcely covered. Current developments in the field of hybrid porous materials hold the potential to make it possible to achieve unequaled FLP-based heterogeneous catalysts with the only limit being the synthetic capacity to prepare their elementary building blocks.

4.3. Towards catalytic systems

From the computations performed, the first and main assumption made was that the FLP was designed to act as a catalyst, i.e. a species used in an understoichiometric amount to allow a chemical reaction to proceed, often faster, and which remains unchanged at the end of the reaction. However, in most of the systems reported, the final product of the hydrogenation of CO2 was an adduct with the Lewis acid or base and the organic hydrogenation target (formic acid, methanol,…) was finally released after hydrolysis. For example, in the work by Zhao42 and Corminbœuf,43 the catalytic activity referred to the moles of formate salt of the protonated Lewis base per mole of Lewis acid. The two studies reported the use of a large excess of Lewis base of 10[thin space (1/6-em)]00042 and 100 equivalents43 compared to the Lewis acid, these systems being however catalytic with respect to the Lewis acid with calculated TONs of 4000 and 24, respectively. As demonstrated by Dyson and Corminbœuf,43 the release of the formate being the limiting step, the use of excess Lewis base is aimed at displacing the equilibrium towards higher CO2 conversions. These pioneering studies underscore the potential for innovative catalytic systems by highlighting the importance of optimizing the ratio of Lewis acids and bases to achieve efficient CO2 hydrogenation. The progress made in understanding and fine-tuning these FLP systems pave the way for future research and development in the field.

4.4. Future developments in computational approaches

There have been a few in-depth DFT studies on solids (periodic calculations or cluster-based simulations) using AIMD and free energy reaction profiles for CO2 hydrogenation. Limitations issues are related to the size of the crystal structure under study. Most studies use UiO-66 or UiO-67 which could be easily handled using their primitive cells. The same approaches might become much more demanding when turning to MOF possessing larger cell sizes and pores (MIL-101, MIL-100 for instance, PCN-777, MOF-808 etc…). Notably, there is currently a blind spot in computational approaches for considering the dynamical behavior of FLPs once grafted or immobilized into the porous host. Such dynamic approaches are however required in order to predict the positioning of the FLPs within the porous host and to restrict the choice of hosts to those allowing the reactivity of the targeted FLPs to be maintained (minimal pore size, confinement effect,161 first coordination sphere interactions with the MOF inorganic sub-network through hydrogen bonding or electrostatics, role of defects). So far, there is no reported theoretical study exploring the participation of the MOF itself in the catalysis as a Lewis partner, typically using open metal sites as acid partners, while facing a grafted basic partner, or vice versa.

The mapping of crossed activity is a significant and achievable milestone. The next frontier is the computational exploration of functionalized MOFs, including the partitioning of multiple LA and LB grafts, as well as their dynamics within porous matrices. With the advanced capabilities of numerous force fields and QM/MM techniques, studying the realistic implantation of such LA/LB pairs is well within reach, promising groundbreaking developments in this field.

5. Conclusions

In conclusion, the last years have seen the tremendous development of molecular FLPs for CO2 hydrogenation, including both experimental and computational advances, with pioneering achievements. Key insights into reaction mechanisms and the factors influencing the reactivity of FLPs towards H2 splitting and subsequent CO2 hydrogenation have been obtained from both experimental and computational studies on molecular FLPs. These have laid the groundwork for developing structure–activity relationships, providing guidelines on how to tune the structural and electronic properties of FLPs to achieve efficient CO2 hydrogenation. Notably, this knowledge can be applied not only to the rational design of molecular FLPs but also to the identification of promising candidates for incorporation into the structure of porous matrices to develop heterogeneous catalysts. However, while the synergy between experimental findings and computations holds the potential to accelerate the discovery of new catalytic FLP systems, examples that leverage the currently available design rules to successfully develop new FLPs for CO2 hydrogenation are still rare. While computational evaluation provides a trend for optimal geometry and reactivity, experimental expertise is key to determine the feasibility of the FLPs’ synthesis and to unravel adequate catalytic conditions.

These challenges become even more complex when transitioning to porous solids through the heterogenization of FLPs. This heterogenization should enable precise control over the FLP's geometry, reactivity, and synergistic effects such as enhancing reactant concentrations locally. Recent pioneering examples already installed a dialogue between computations and experimentations but the implementation of computationally designed in experimentally constructed FLP still require intensive efforts for their in catalytic applications. The development of computational methodologies to go even deeper into the predictions of the synthetic and catalytic conditions combined with the proliferation of synthesis methods based on almost infinite variations of compositions and structures of porous hybrid solids is aimed to push the limits of current systems. Given the evolving nature of this research field, the collaborative efforts of synthetic and computational chemistry are poised to yield sustainable FLP-based catalytic systems for direct CO2 hydrogenation in the foreseeable future.

Data availability

No primary research results, software or code have been included and no new data were generated or analysed as part of this review.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the French National Agency as part of the program No. ANR-21-CE07-0028 (FLIPS). A.S.-D. also acknowledges the Spanish Ministry of Universities and the European Union – Next Generation EU for financial support through a Margarita Salas grant.

References

  1. S. Hübner, J. G. de Vries and V. Farina, Why Does Industry Not Use Immobilized Transition Metal Complexes as Catalysts, Adv. Synth. Catal., 2016, 358(1), 3–25,  DOI:10.1002/adsc.201500846 .
  2. J. Artz, T. E. Müller, K. Thenert, J. Kleinekorte, R. Meys, A. Sternberg, A. Bardow and W. Leitner, Sustainable Conversion of Carbon Dioxide: An Integrated Review of Catalysis and Life Cycle Assessment, Chem. Rev., 2018, 118(2), 434–504,  DOI:10.1021/acs.chemrev.7b00435 .
  3. T. Gaich and P. S. Baran, Aiming for the Ideal Synthesis, J. Org. Chem., 2010, 75(14), 4657–4673,  DOI:10.1021/jo1006812 .
  4. G. C. Welch, R. R. S. Juan, J. D. Masuda and D. W. Stephan, Reversible, Metal-Free Hydrogen Activation, Science, 2006, 314(5802), 1124–1126,  DOI:10.1126/science.1134230 .
  5. M. Navarro; J. J. Moreno and J. Campos, Frustrated Lewis Pair Systems, in Comprehensive Organometallic Chemistry IV, 2022, vol. 10, pp. 523–616 Search PubMed .
  6. J. Paradies, From Structure to Novel Reactivity in Frustrated Lewis Pairs, Coord. Chem. Rev., 2019, 380, 170–183,  DOI:10.1016/j.ccr.2018.09.014 .
  7. B. L. Frenette and E. Rivard, Frustrated Lewis Pair Chelation in the P-Block, Chem. – Eur. J., 2023, 29(65), e202302332,  DOI:10.1002/chem.202302332 .
  8. D. W. Stephan and G. Erker, Frustrated Lewis Pair Chemistry of Carbon, Nitrogen and Sulfur Oxides, Chem. Sci., 2014, 5(7), 2625–2641,  10.1039/C4SC00395K .
  9. D. W. Stephan, Frustrated Lewis Pairs: From Concept to Catalysis, Acc. Chem. Res., 2015, 48(2), 306–316,  DOI:10.1021/ar500375j .
  10. M. N. Khan, Y. Ingen, T. van; Boruah, A. McLauchlan, T. Wirth and R. L. Melen, Advances in CO2 Activation by Frustrated Lewis Pairs: From Stoichiometric to Catalytic Reactions, Chem. Sci., 2023, 14(47), 13661–13695,  10.1039/D3SC03907B .
  11. M. Pérez-Jiménez, H. Corona, F. de la Cruz-Martínez and J. Campos, Donor-Acceptor Activation of Carbon Dioxide, Chem. – Eur. J., 2023, 29(61), e202301428,  DOI:10.1002/chem.202301428 .
  12. D. J. Scott, M. J. Fuchter and A. E. Ashley, Designing Effective ‘Frustrated Lewis Pair’ Hydrogenation Catalysts, Chem. Soc. Rev., 2017, 46(19), 5689–5700,  10.1039/C7CS00154A .
  13. F.-G. Fontaine, M.-A. Courtemanche, M.-A. Légaré and É. Rochette, Design Principles in Frustrated Lewis Pair Catalysis for the Functionalization of Carbon Dioxide and Heterocycles, Coord. Chem. Rev., 2017, 334, 124–135,  DOI:10.1016/j.ccr.2016.05.005 .
  14. A. R. Jupp and D. W. Stephan, New Directions for Frustrated Lewis Pair Chemistry, Trends Chem., 2019, 1(1), 35–48,  DOI:10.1016/j.trechm.2019.01.006 .
  15. Y. Ma, S. Zhang, C.-R. Chang, Z.-Q. Huang, J. C. Ho and Y. Qu, Semi-Solid and Solid Frustrated Lewis Pair Catalysts, Chem. Soc. Rev., 2018, 47(15), 5541–5553,  10.1039/C7CS00691H .
  16. M. Burai Patrascu, J. Pottel, S. Pinus, M. Bezanson, P.-O. Norrby and N. Moitessier, From Desktop to Benchtop with Automated Computational Workflows for Computer-Aided Design in Asymmetric Catalysis, Nat. Catal., 2020, 3(7), 574–584,  DOI:10.1038/s41929-020-0468-3 .
  17. A. R. Rosales, J. Wahlers, E. Limé, R. E. Meadows, K. W. Leslie, R. Savin, F. Bell, E. Hansen, P. Helquist, R. H. Munday, O. Wiest and P.-O. Norrby, Rapid Virtual Screening of Enantioselective Catalysts Using CatVS, Nat. Catal., 2019, 2(1), 41–45,  DOI:10.1038/s41929-018-0193-3 .
  18. A. R. Rosales, T. R. Quinn, J. Wahlers, A. Tomberg, X. Zhang, P. Helquist, O. Wiest and P.-O. Norrby, Application of Q2MM to Predictions in Stereoselective Synthesis, Chem. Commun., 2018, 54(60), 8294–8311,  10.1039/C8CC03695K .
  19. P. Mialane, C. Mellot-Draznieks, P. Gairola, M. Duguet, Y. Benseghir, O. Oms and A. Dolbecq, Heterogenisation of Polyoxometalates and Other Metal-Based Complexes in Metal–Organic Frameworks: From Synthesis to Characterisation and Applications in Catalysis, Chem. Soc. Rev., 2021, 50(10), 6152–6220,  10.1039/D0CS00323A .
  20. B. W. J. Chen, L. Xu and M. Mavrikakis, Computational Methods in Heterogeneous Catalysis, Chem. Rev., 2021, 121(2), 1007–1048,  DOI:10.1021/acs.chemrev.0c01060 .
  21. X. Shen, Z. Wang, X. J. Gao and X. Gao, Reaction Mechanisms and Kinetics of Nanozymes: Insights from Theory and Computation, Adv. Mater., 2024, 36(10), 2211151,  DOI:10.1002/adma.202211151 .
  22. P. Sit and L. Zhang, Density Functional Theory in Heterogeneous Catalysis, in Heterogeneous Catalysts, John Wiley & Sons, Ltd, 2021, pp. 405–418 DOI:10.1002/9783527813599.ch23 .
  23. D. W. Stephan, Catalysis, FLPs, and Beyond, Chem, 2020, 6(7), 1520–1526,  DOI:10.1016/j.chempr.2020.05.007 .
  24. P. Sreejyothi and S. K. Mandal, From CO2 Activation to Catalytic Reduction: A Metal-Free Approach, Chem. Sci., 2020, 11(39), 10571–10593,  10.1039/D0SC03528A .
  25. N. Li and W. Zhang, Frustrated Lewis Pairs: Discovery and Overviews in Catalysis, Chin. J. Chem., 2020, 38(11), 1360–1370,  DOI:10.1002/cjoc.202000027 .
  26. J. Lam, K. M. Szkop, E. Mosaferi and D. W. Stephan, FLP Catalysis: Main Group Hydrogenations of Organic Unsaturated Substrates, Chem. Soc. Rev., 2019, 48(13), 3592–3612,  10.1039/C8CS00277K .
  27. M. Chen, M. Wang and M. Ji, Design and Synthesis of Heterogeneous Frustrated Lewis Pairs for Hydrogenation: From Molecular Immobilization to Defects Engineering, ChemCatChem, 2023, e202301472,  DOI:10.1002/cctc.202301472 .
  28. Y. Zhang, P. C. Lan, K. Martin and S. Ma, Porous Frustrated Lewis Pair Catalysts: Advances and Perspective, Chem Catal., 2022, 2(3), 439–457,  DOI:10.1016/j.checat.2021.12.001 .
  29. G. Sharma, P. D. Newman and J. A. Platts, A Review of Quantum Chemical Studies of Frustrated Lewis Pairs, J. Mol. Graphics Modell., 2021, 105, 107846,  DOI:10.1016/j.jmgm.2021.107846 .
  30. F.-G. Fontaine and É. Rochette, Ambiphilic Molecules: From Organometallic Curiosity to Metal-Free Catalysts, Acc. Chem. Res., 2018, 51(2), 454–464,  DOI:10.1021/acs.accounts.7b00514 .
  31. A. E. Ashley, A. L. Thompson and D. O’Hare, Non-Metal-Mediated Homogeneous Hydrogenation of CO2 to CH3OH, Angew. Chem., Int. Ed., 2009, 48(52), 9839–9843,  DOI:10.1002/anie.200905466 .
  32. M.-A. Courtemanche, A. P. Pulis, É. Rochette, M.-A. Légaré, D. W. Stephan and F.-G. Fontaine, Intramolecular B/N Frustrated Lewis Pairs and the Hydrogenation of Carbon Dioxide, Chem. Commun., 2015, 51(48), 9797–9800,  10.1039/C5CC03072B .
  33. F.-G. Fontaine and D. W. Stephan, Metal-Free Reduction of CO2, Curr. Opin. Green Sustainable Chem., 2017, 3, 28–32,  DOI:10.1016/j.cogsc.2016.11.004 .
  34. S. D. Tran, T. A. Tronic, W. Kaminsky, D. Michael Heinekey and J. M. Mayer, Metal-Free Carbon Dioxide Reduction and Acidic C–H Activations Using a Frustrated Lewis Pair, Inorg. Chim. Acta, 2011, 369(1), 126–132,  DOI:10.1016/j.ica.2010.12.022 .
  35. T. Voss, T. Mahdi, E. Otten, R. Fröhlich, G. Kehr, D. W. Stephan and G. Erker, Frustrated Lewis Pair Behavior of Intermolecular Amine/B(C6F5)3 Pairs, Organometallics, 2012, 31(6), 2367–2378,  DOI:10.1021/om300017u .
  36. C. M. Mömming, E. Otten, G. Kehr, R. Fröhlich, S. Grimme, D. W. Stephan and G. Erker, Reversible Metal-Free Carbon Dioxide Binding by Frustrated Lewis Pairs, Angew. Chem., Int. Ed., 2009, 48(36), 6643–6646,  DOI:10.1002/anie.200901636 .
  37. A. L. Travis, S. C. Binding, H. Zaher, T. A. Q. Arnold, J.-C. Buffet and D. O’Hare, Small Molecule Activation by Frustrated Lewis Pairs, Dalton Trans., 2013, 42(7), 2431–2437,  10.1039/C2DT32525J .
  38. G. Ménard and D. W. Stephan, Room Temperature Reduction of CO2 to Methanol by Al-Based Frustrated Lewis Pairs and Ammonia Borane, J. Am. Chem. Soc., 2010, 132(6), 1796–1797,  DOI:10.1021/ja9104792 .
  39. G. Ménard and D. W. Stephan, Stoichiometric Reduction of CO2 to CO by Aluminum-Based Frustrated Lewis Pairs, Angew. Chem., Int. Ed., 2011, 50(36), 8396–8399,  DOI:10.1002/anie.201103600 .
  40. C. M. Zall, J. C. Linehan and A. M. Appel, A Molecular Copper Catalyst for Hydrogenation of CO2 to Formate, ACS Catal., 2015, 5(9), 5301–5305,  DOI:10.1021/acscatal.5b01646 .
  41. L. Wickemeyer, N. Aders, A. Mix, B. Neumann, H.-G. Stammler, J. J. Cabrera-Trujillo, I. Fernández and N. W. Mitzel, Carbon Dioxide Reduction by an Al–O–P Frustrated Lewis Pair, Chem. Sci., 2022, 13(27), 8088–8094,  10.1039/D2SC01870E .
  42. T. Zhao, X. Hu, Y. Wu and Z. Zhang, Hydrogenation of CO2 to Formate with H2: Transition Metal Free Catalyst Based on a Lewis Pair, Angew. Chem., Int. Ed., 2019, 58(3), 722–726,  DOI:10.1002/anie.201809634 .
  43. S. Das, R. C. Turnell-Ritson, P. J. Dyson and C. Corminboeuf, Design of Frustrated Lewis Pair Catalysts for Direct Hydrogenation of CO2, Angew. Chem., Int. Ed., 2022, 61(46), e202208987,  DOI:10.1002/anie.202208987 .
  44. A. G. Massey, A. J. Park and F. G. A. Stone, Tris(Pentafluorophenyl)Boron, Proc. Chem. Soc., 1963, 212–213,  DOI:10.1002/anie.201704097 .
  45. A. G. Massey and A. J. Park, Perfluorophenyl Derivatives of the Elements: I. Tris(Pentafluorophenyl)Boron, J. Organomet. Chem., 1964, 2(3), 245–250,  DOI:10.1016/S0022-328X(00)80518-5 .
  46. W. E. Piers and T. Chivers, Pentafluorophenylboranes: From Obscurity to Applications, Chem. Soc. Rev., 1997, 26(5), 345–354,  10.1039/CS9972600345 .
  47. G. Erker, Tris(Pentafluorophenyl)Borane: A Special Boron Lewis Acid for Special Reactions, Dalton Trans., 2005, 1883–1890,  10.1039/B503688G .
  48. J. L. Carden, A. Dasgupta and R. L. Melen, Halogenated Triarylboranes: Synthesis, Properties and Applications in Catalysis, Chem. Soc. Rev., 2020, 49(6), 1706–1725,  10.1039/C9CS00769E .
  49. V. Sumerin, F. Schulz, M. Nieger, M. Leskelä, T. Repo and B. Rieger, Facile Heterolytic H2 Activation by Amines and B(C6F5)3, Angew. Chem., Int. Ed., 2008, 47(32), 6001–6003,  DOI:10.1002/anie.200800935 .
  50. L. Liu, N. Vankova and T. Heine, A Kinetic Study on the Reduction of CO2 by Frustrated Lewis Pairs: From Understanding to Rational Design, Phys. Chem. Chem. Phys., 2016, 18(5), 3567–3574,  10.1039/C5CP06925D .
  51. H. Wang, Y. Zhao, H. Zhao, J. Yang, D. Zhai, L. Sun and W. Deng, In Silico Design of Metal-Free Hydrophosphate Catalysts for Hydrogenation of CO2 to Formate, Phys. Chem. Chem. Phys., 2022, 24(5), 2901–2908,  10.1039/D1CP04582B .
  52. S. J. Geier and D. W. Stephan, Lutidine/B(C6F5)3: At the Boundary of Classical and Frustrated Lewis Pair Reactivity, J. Am. Chem. Soc., 2009, 131(10), 3476–3477,  DOI:10.1021/ja900572x .
  53. T. A. Rokob, A. Hamza and I. Pápai, Rationalizing the Reactivity of Frustrated Lewis Pairs: Thermodynamics of H2 Activation and the Role of Acid–Base Properties, J. Am. Chem. Soc., 2009, 131(30), 10701–10710,  DOI:10.1021/ja903878z .
  54. I. Peuser, R. C. Neu, X. Zhao, M. Ulrich, B. Schirmer, J. A. Tannert, G. Kehr, R. Fröhlich, S. Grimme, G. Erker and D. W. Stephan, CO2 and Formate Complexes of Phosphine/Borane Frustrated Lewis Pairs, Chem. – Eur. J., 2011, 17(35), 9640–9650,  DOI:10.1002/chem.201100286 .
  55. G. C. Welch and D. W. Stephan, Facile Heterolytic Cleavage of Dihydrogen by Phosphines and Boranes, J. Am. Chem. Soc., 2007, 129(7), 1880–1881,  DOI:10.1021/ja067961j .
  56. S. C. Binding, H. Zaher, F. M. Chadwick and D. O’Hare, Heterolytic Activation of Hydrogen Using Frustrated Lewis Pairs Containing Tris(2,2′,2′′-Perfluorobiphenyl)Borane, Dalton Trans., 2012, 41(30), 9061–9066,  10.1039/C2DT30334E .
  57. X. Zhao and D. W. Stephan, Bis-Boranes in the Frustrated Lewis Pair Activation of Carbon Dioxide, Chem. Commun., 2011, 47(6), 1833–1835,  10.1039/C0CC04791K .
  58. M. J. Sgro, J. Dömer and D. W. Stephan, Stoichiometric CO2 Reductions Using a Bis-Borane-Based Frustrated Lewis Pair, Chem. Commun., 2012, 48(58), 7253–7255,  10.1039/C2CC33301E .
  59. S. A. Weicker and D. W. Stephan, Activation of Carbon Dioxide by Silyl Triflate-Based Frustrated Lewis Pairs, Chem. – Eur. J., 2015, 21(37), 13027–13034,  DOI:10.1002/chem.201501904 .
  60. D. W. Stephan, Frustrated Lewis Pairs, J. Am. Chem. Soc., 2015, 137(32), 10018–10032,  DOI:10.1021/jacs.5b06794 .
  61. Y. Jiang, O. Blacque, T. Fox and H. Berke, 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(20), 7751–7760,  DOI:10.1021/ja402381d .
  62. M.-A. Courtemanche, M.-A. Légaré, L. Maron and F.-G. Fontaine, A Highly Active Phosphine–Borane Organocatalyst for the Reduction of CO2 to Methanol Using Hydroboranes, J. Am. Chem. Soc., 2013, 135(25), 9326–9329,  DOI:10.1021/ja404585p .
  63. M.-A. Courtemanche, M.-A. Légaré, L. Maron and F.-G. Fontaine, Reducing CO2 to Methanol Using Frustrated Lewis Pairs: On the Mechanism of Phosphine–Borane-Mediated Hydroboration of CO2, J. Am. Chem. Soc., 2014, 136(30), 10708–10717,  DOI:10.1021/ja5047846 .
  64. M.-A. Courtemanche, J. Larouche, M.-A. Légaré, W. Bi, L. Maron and F.-G. Fontaine, A Tris(Triphenylphosphine)Aluminum Ambiphilic Precatalyst for the Reduction of Carbon Dioxide with Catecholborane, Organometallics, 2013, 32(22), 6804–6811,  DOI:10.1021/om400645s .
  65. R. Declercq, G. Bouhadir, D. Bourissou, M.-A. Légaré, M.-A. Courtemanche, K. S. Nahi, N. Bouchard, F.-G. Fontaine and L. Maron, Hydroboration of Carbon Dioxide Using Ambiphilic Phosphine–Borane Catalysts: On the Role of the Formaldehyde Adduct, ACS Catal., 2015, 5(4), 2513–2520,  DOI:10.1021/acscatal.5b00189 .
  66. F. Krämer, J. Paradies, I. Fernández and F. Breher, Quo Vadis CO2 Activation: Catalytic Reduction of CO2 to Methanol Using Aluminum and Gallium/Carbon-Based Ambiphiles, Chem. – Eur. J., 2024, 30(5), e202303380,  DOI:10.1002/chem.202303380 .
  67. T. Wang and D. W. Stephan, Phosphine Catalyzed Reduction of CO2 with Boranes, Chem. Commun., 2014, 50(53), 7007–7010,  10.1039/C4CC02103G .
  68. S. Das, R. Laplaza, J. T. Blaskovits and C. Corminboeuf, Mapping Active Site Geometry to Activity in Immobilized Frustrated Lewis Pair Catalysts, Angew. Chem., Int. Ed., 2022, 61(32), e202202727,  DOI:10.1002/anie.202202727 .
  69. K. C. Szeto, W. Sahyoun, N. Merle, J. L. Castelbou, N. Popoff, F. Lefebvre, J. Raynaud, C. Godard, C. Claver, L. Delevoye, R. M. Gauvin and M. Taoufik, Development of Silica-Supported Frustrated Lewis Pairs: Highly Active Transition Metal-Free Catalysts for the Z-Selective Reduction of Alkynes, Catal. Sci. Technol., 2016, 6(3), 882–889,  10.1039/C5CY01372K .
  70. J.-Y. Xing, J.-C. Buffet, N. H. Rees, P. Nørby and D. O’Hare, Hydrogen Cleavage by Solid-Phase Frustrated Lewis Pairs, Chem. Commun., 2016, 52(69), 10478–10481,  10.1039/C6CC04937K .
  71. K. Mentoor, L. Twigge, J. W. H. Niemantsverdriet, J. C. Swarts and E. Erasmus, Silica Nanopowder Supported Frustrated Lewis Pairs for CO2 Capture and Conversion to Formic Acid, Inorg. Chem., 2021, 60(1), 55–69,  DOI:10.1021/acs.inorgchem.0c02012 .
  72. Z. Niu, W. Zhang, P. C. Lan, B. Aguila and S. Ma, Promoting Frustrated Lewis Pairs for Heterogeneous Chemoselective Hydrogenation via the Tailored Pore Environment within Metal–Organic Frameworks, Angew. Chem., Int. Ed., 2019, 58(22), 7420–7424,  DOI:10.1002/anie.201903763 .
  73. Z. Niu, W. D. C. Bhagya Gunatilleke, Q. Sun, P. C. Lan, J. Perman, J.-G. Ma, Y. Cheng, B. Aguila and S. Ma, Metal-Organic Framework Anchored with a Lewis Pair as a New Paradigm for Catalysis, Chem, 2018, 4(11), 2587–2599,  DOI:10.1016/j.chempr.2018.08.018 .
  74. S. Shyshkanov, T. N. Nguyen, F. M. Ebrahim, K. C. Stylianou and P. J. Dyson, 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(16), 5371–5375,  DOI:10.1002/anie.201901171 .
  75. S. Shyshkanov, T. N. Nguyen, A. Chidambaram, K. C. Stylianou and P. J. Dyson, Frustrated Lewis Pair-Mediated Fixation of CO2 within a Metal–Organic Framework, Chem. Commun., 2019, 55(73), 10964–10967,  10.1039/C9CC04374H .
  76. Y. Zhang, S. Chen, A. M. Al-Enizi, A. Nafady, Z. Tang and S. Ma, Chiral Frustrated Lewis Pair@Metal-Organic Framework as a New Platform for Heterogeneous Asymmetric Hydrogenation, Angew. Chem., Int. Ed., 2023, 62(2), e202213399,  DOI:10.1002/anie.202213399 .
  77. M. Trunk, J. F. Teichert and A. Thomas, Room-Temperature Activation of Hydrogen by Semi-Immobilized Frustrated Lewis Pairs in Microporous Polymer Networks, J. Am. Chem. Soc., 2017, 139(10), 3615–3618,  DOI:10.1021/jacs.6b13147 .
  78. A. Willms, H. Schumacher, T. Tabassum, L. Qi, S. L. Scott, P. J. C. Hausoul and M. Rose, Solid Molecular Frustrated Lewis Pairs in a Polyamine Organic Framework for the Catalytic Metal-Free Hydrogenation of Alkenes, ChemCatChem, 2018, 10(8), 1835–1843,  DOI:10.1002/cctc.201701783 .
  79. L. Chen, R. Liu and Q. Yan, Polymer Meets Frustrated Lewis Pair: Second-Generation CO2-Responsive Nanosystem for Sustainable CO2 Conversion, Angew. Chem., Int. Ed., 2018, 57(30), 9336–9340,  DOI:10.1002/anie.201804034 .
  80. L. Hou, Y. Liang, Q. Wang, Y. Zhang, D. Dong and N. Zhang, Lewis Pair-Mediated Surface-Initiated Polymerization, ACS Macro Lett., 2018, 7(1), 65–69,  DOI:10.1021/acsmacrolett.7b00903 .
  81. J. W. Maina, C. Pozo-Gonzalo, L. Kong, J. Schütz, M. Hill and L. F. Dumée, Metal Organic Framework Based Catalysts for CO2 Conversion, Mater. Horiz., 2017, 4(3), 345–361,  10.1039/C6MH00484A .
  82. J. Canivet and F. M. Wisser, Metal–Organic Framework Catalysts for Solar Fuels: Light-Driven Conversion of Carbon Dioxide into Formic Acid, ACS Appl. Energy Mater., 2023, 6(18), 9027–9043,  DOI:10.1021/acsaem.2c03731 .
  83. H. Beyzavi, R. C. Klet, S. Tussupbayev, J. Borycz, N. A. Vermeulen, C. J. Cramer, J. F. Stoddart, J. T. Hupp and O. K. Farha, A Hafnium-Based Metal–Organic Framework as an Efficient and Multifunctional Catalyst for Facile CO2 Fixation and Regioselective and Enantioretentive Epoxide Activation, J. Am. Chem. Soc., 2014, 136(45), 15861–15864,  DOI:10.1021/ja508626n .
  84. W.-Y. Gao, Y. Chen, Y. Niu, K. Williams, L. Cash, P. J. Perez, L. Wojtas, J. Cai, Y.-S. Chen and S. Ma, Crystal Engineering of an Nbo Topology Metal–Organic Framework for Chemical Fixation of CO2 under Ambient Conditions, Angew. Chem., Int. Ed., 2014, 53(10), 2615–2619,  DOI:10.1002/anie.201309778 .
  85. K. Sumida, D. L. Rogow, J. A. Mason, T. M. McDonald, E. D. Bloch, Z. R. Herm, T.-H. Bae and J. R. Long, Carbon Dioxide Capture in Metal–Organic Frameworks, Chem. Rev., 2012, 112(2), 724–781,  DOI:10.1021/cr2003272 .
  86. Z. Ji, H. Wang, S. Canossa, S. Wuttke and O. M. Yaghi, Pore Chemistry of Metal–Organic Frameworks, Adv. Funct. Mater., 2020, 30(41), 2000238,  DOI:10.1002/adfm.202000238 .
  87. A. Dutta, Y. Pan, J.-Q. Liu and A. Kumar, Multicomponent Isoreticular Metal-Organic Frameworks: Principles, Current Status and Challenges, Coord. Chem. Rev., 2021, 445, 214074,  DOI:10.1016/j.ccr.2021.214074 .
  88. J. Ye and J. K. Johnson, Design of Lewis Pair-Functionalized Metal Organic Frameworks for CO2 Hydrogenation, ACS Catal., 2015, 5(5), 2921–2928,  DOI:10.1021/acscatal.5b00396 .
  89. J. Ye and J. K. Johnson, Screening Lewis Pair Moieties for Catalytic Hydrogenation of CO2 in Functionalized UiO-66, ACS Catal., 2015, 5(10), 6219–6229,  DOI:10.1021/acscatal.5b01191 .
  90. J. Ye and J. K. Johnson, Catalytic Hydrogenation of CO2 to Methanol in a Lewis Pair Functionalized MOF, Catal. Sci. Technol., 2016, 6(24), 8392–8405,  10.1039/C6CY01245K .
  91. J. Ye, B. Y. Yeh, R. A. Reynolds and J. K. Johnson, Screening the Activity of Lewis Pairs for Hydrogenation of CO2, Mol. Simul., 2017, 43(10–11), 821–827,  DOI:10.1080/08927022.2017.1295457 .
  92. J. Ye, L. Li and J. K. Johnson, The Effect of Topology in Lewis Pair Functionalized Metal Organic Frameworks on CO2 Adsorption and Hydrogenation, Catal. Sci. Technol., 2018, 8(18), 4609–4617,  10.1039/C8CY01018H .
  93. M. Heshmat, Alternative Pathway of CO2 Hydrogenation by Lewis-Pair-Functionalized UiO-66 MOF Revealed by Metadynamics Simulations, J. Phys. Chem. C, 2020, 124(20), 10951–10960,  DOI:10.1021/acs.jpcc.0c01088 .
  94. G. Férey, C. Mellot-Draznieks, C. Serre, F. Millange, J. Dutour, S. Surblé and I. Margiolaki, A Chromium Terephthalate-Based Solid with Unusually Large Pore Volumes and Surface Area, Science, 2005, 309(5743), 2040–2042,  DOI:10.1126/science.1116275 .
  95. Q. Liu, Q. Liao, J. Hu, K. Xi, Y. Wu and X. Hu, Covalent Organic Frameworks Anchored with Frustrated Lewis Pairs for Hydrogenation of Alkynes with H2, J. Mater. Chem. A, 2022, 10(13), 7333–7340,  10.1039/D1TA08916A .
  96. M. V. Zakharova, N. Masoumifard, Y. Hu, J. Han, F. Kleitz and F.-G. Fontaine, Designed Synthesis of Mesoporous Solid-Supported Lewis Acid–Base Pairs and Their CO2 Adsorption Behaviors, ACS Appl. Mater. Interfaces, 2018, 10(15), 13199–13210,  DOI:10.1021/acsami.8b00640 .
  97. D. Yepes, P. Jaque and I. Fernández, Deeper Insight into the Factors Controlling H2 Activation by Geminal Aminoborane-Based Frustrated Lewis Pairs, Chem. – Eur. J., 2016, 22(52), 18801–18809,  DOI:10.1002/chem.201603889 .
  98. L. Liu Zeonjuk, P. St Petkov, T. Heine, G.-V. Röschenthaler, J. Eicher and N. Vankova, Are Intramolecular Frustrated Lewis Pairs Also Intramolecular Catalysts? A Theoretical Study on H2 Activation, Phys. Chem. Chem. Phys., 2015, 17(16), 10687–10698,  10.1039/C5CP00368G .
  99. T. A. Rokob, A. Hamza, A. Stirling, T. Soós and I. Pápai, Turning Frustration into Bond Activation: A Theoretical Mechanistic Study on Heterolytic Hydrogen Splitting by Frustrated Lewis Pairs, Angew. Chem., Int. Ed., 2008, 47(13), 2435–2438,  DOI:10.1002/anie.200705586 .
  100. Y. Guo and S. Li, Unusual Concerted Lewis Acid–Lewis Base Mechanism for Hydrogen Activation by a Phosphine–Borane Compound, Inorg. Chem., 2008, 47(14), 6212–6219,  DOI:10.1021/ic702489s .
  101. P. Spies, G. Erker, G. Kehr, K. Bergander, R. Fröhlich, S. Grimme and D. W. Stephan, Rapid Intramolecular Heterolytic Dihydrogen Activation by a Four-Membered Heterocyclic Phosphane–Borane Adduct, Chem. Commun., 2007, 5072,  10.1039/b710475h .
  102. P. Spies, S. Schwendemann, S. Lange, G. Kehr, R. Fröhlich and G. Erker, Metal-Free Catalytic Hydrogenation of Enamines, Imines, and Conjugated Phosphinoalkenylboranes, Angew. Chem., Int. Ed., 2008, 47(39), 7543–7546,  DOI:10.1002/anie.200801432 .
  103. P. Spies, G. Kehr, K. Bergander, B. Wibbeling, R. Fröhlich and G. Erker, Metal-Free Dihydrogen Activation Chemistry: Structural and Dynamic Features of Intramolecular P/B Pairs, Dalton Trans., 2009, 1534–1541,  10.1039/b815832k .
  104. S. Grimme, H. Kruse, L. Goerigk and G. Erker, The Mechanism of Dihydrogen Activation by Frustrated Lewis Pairs Revisited, Angew. Chem., Int. Ed., 2010, 49(8), 1402–1405,  DOI:10.1002/anie.200905484 .
  105. T. A. Rokob, I. Bakó, A. Stirling, A. Hamza and I. Pápai, Reactivity Models of Hydrogen Activation by Frustrated Lewis Pairs: Synergistic Electron Transfers or Polarization by Electric Field, J. Am. Chem. Soc., 2013, 135(11), 4425–4437,  DOI:10.1021/ja312387q .
  106. J. J. Cabrera-Trujillo and I. Fernández, Aromaticity Can Enhance the Reactivity of P-Donor/Borole Frustrated Lewis Pairs, Chem. Commun., 2019, 55(5), 675–678,  10.1039/C8CC09777A .
  107. M. Pu and T. Privalov, Ab Initio Dynamics Trajectory Study of the Heterolytic Cleavage of H2 by a Lewis Acid [B(C6F5)3] and a Lewis Base [P(tBu)3], J. Chem. Phys., 2013, 138(15), 154305,  DOI:10.1063/1.4799932 .
  108. M. Pu and T. Privalov, How Frustrated Lewis Acid/Base Systems Pass through Transition-State Regions: H2 Cleavage by [tBu3P/B(C6F5)3], Chem. Phys. Chem., 2014, 15(14), 2936–2944,  DOI:10.1002/cphc.201402450 .
  109. M. Pu and T. Privalov, Ab Initio Molecular Dynamics Study of Hydrogen Cleavage by a Lewis Base [tBu3 P] and a Lewis Acid [B(C6F5)3] at the Mesoscopic Level-Dynamics in the Solute-Solvent Molecular Clusters, Chem. Phys. Chem., 2014, 15(17), 3714–3719,  DOI:10.1002/cphc.201402519 .
  110. M. Pu and T. Privalov, Chemistry of Intermolecular Frustrated Lewis Pairs in Motion: Emerging Perspectives and Prospects, Isr. J. Chem., 2015, 55(2), 179–195,  DOI:10.1002/ijch.201400159 .
  111. L. Liu, B. Lukose and B. Ensing, Hydrogen Activation by Frustrated Lewis Pairs Revisited by Metadynamics Simulations, J. Phys. Chem. C, 2017, 121(4), 2046–2051,  DOI:10.1021/acs.jpcc.6b09991 .
  112. M. Heshmat and T. Privalov, Testing the Nature of Reaction Coordinate Describing Interaction of H2 with Carbonyl Carbon, Activated by Lewis Acid Complexation, and the Lewis Basic Solvent: A Born-Oppenheimer Molecular Dynamics Study with Explicit Solvent, J. Chem. Phys., 2017, 147(9), 094302,  DOI:10.1063/1.4999708 .
  113. G. Skara, F. De Vleeschouwer, P. Geerlings, F. De Proft and B. Pinter, Heterolytic Splitting of Molecular Hydrogen by Frustrated and Classical Lewis Pairs: A Unified Reactivity Concept, Sci. Rep., 2017, 7(1), 16024,  DOI:10.1038/s41598-017-16244-1 .
  114. M. Heshmat and B. Ensing, Optimizing the Energetics of FLP-Type H2 Activation by Modulating the Electronic and Structural Properties of the Lewis Acids: A DFT Study, J. Phys. Chem. A, 2020, 124(32), 6399–6410,  DOI:10.1021/acs.jpca.0c03108 .
  115. J. Daru, I. Bakó, A. Stirling and I. Pápai, Mechanism of Heterolytic Hydrogen Splitting by Frustrated Lewis Pairs: Comparison of Static and Dynamic Models, ACS Catal., 2019, 9(7), 6049–6057,  DOI:10.1021/acscatal.9b01137 .
  116. A. Hamza, A. Stirling, T. András Rokob and I. Pápai, Mechanism of Hydrogen Activation by Frustrated Lewis Pairs: A Molecular Orbital Approach, Int. J. Quantum Chem., 2009, 109(11), 2416–2425,  DOI:10.1002/qua.22203 .
  117. M. Ullrich, A. J. Lough and D. W. Stephan, Reversible, Metal-Free, Heterolytic Activation of H2 at Room Temperature, J. Am. Chem. Soc., 2009, 131(1), 52–53,  DOI:10.1021/ja808506t .
  118. L. Liu, B. Lukose, P. Jaque and B. Ensing, Reaction Mechanism of Hydrogen Activation by Frustrated Lewis Pairs, Green Energy Environ., 2019, 4(1), 20–28,  DOI:10.1016/j.gee.2018.06.001 .
  119. D. W. Stephan and G. Erker, Frustrated Lewis Pairs: Metal-Free Hydrogen Activation and More, Angew. Chem., Int. Ed., 2010, 49(1), 46–76,  DOI:10.1002/anie.200903708 .
  120. P. Pérez, D. Yepes, P. Jaque, E. Chamorro, L. R. Domingo, R. S. Rojas and A. Toro-Labbé, A Computational and Conceptual DFT Study on the Mechanism of Hydrogen Activation by Novel Frustrated Lewis Pairs, Phys. Chem. Chem. Phys., 2015, 17(16), 10715–10725,  10.1039/C5CP00306G .
  121. D. Zhuang, Y. Li and J. Zhu, Antiaromaticity-Promoted Activation of Dihydrogen with Borole Fused Cyclooctatetraene Frustrated Lewis Pairs: A Density Functional Theory Study, Organometallics, 2020, 39(14), 2636–2641,  DOI:10.1021/acs.organomet.0c00263 .
  122. D. M. Camaioni, B. Ginovska-Pangovska, G. K. Schenter, S. M. Kathmann and T. Autrey, Analysis of the Activation and Heterolytic Dissociation of H2 by Frustrated Lewis Pairs: NH 3/BX3 (X = H, F, and Cl), J. Phys. Chem. A, 2012, 116(26), 7228–7237,  DOI:10.1021/jp3039829 .
  123. J. Zhang, Y. Shao, Y. Li, Y. Liu and Z. Ke, Rational Design of FLP Catalysts for Reversible H2 Activation: A DFT Study of the Geometric and Electronic Effects, Chin. Chem. Lett., 2018, 29(8), 1226–1232,  DOI:10.1016/j.cclet.2018.02.007 .
  124. E. L. Kolychev; E. Theuergarten and M. Tamm, N-Heterocyclic Carbenes in FLP Chemistry, in Frustrated Lewis Pairs II, Topics in Current Chemistry, ed. G. Erker and D. W. Stephan, Springer Berlin Heidelberg, Berlin, Heidelberg, 2012, vol. 334, pp. 121–155 DOI:10.1007/128_2012_379 .
  125. D. F. Wass and A. M. Chapman, Frustrated Lewis Pairs Beyond the Main Group: Transition Metal-Containing Systems, in Frustrated Lewis Pairs II, Topics in Current Chemistry, ed. G. Erker and D. W. Stephan, Springer Berlin Heidelberg, Berlin, Heidelberg, 2013, vol. 334, pp. 261–280 DOI:10.1007/128_2012_395 .
  126. Q. Wan, S. Lin and H. Guo, Frustrated Lewis Pairs in Heterogeneous Catalysis: Theoretical Insights, Molecules, 2022, 27(12), 3734,  DOI:10.3390/molecules27123734 .
  127. K. K. Ghuman, L. B. Hoch, T. E. Wood, C. Mims, C. V. Singh and G. A. Ozin, Surface Analogues of Molecular Frustrated Lewis Pairs in Heterogeneous CO2 Hydrogenation Catalysis, ACS Catal., 2016, 6(9), 5764–5770,  DOI:10.1021/acscatal.6b01015 .
  128. M. Ghoussoub, S. Yadav, K. K. Ghuman, G. A. Ozin and C. V. Singh, Metadynamics-Biased Ab Initio Molecular Dynamics Study of Heterogeneous CO2 Reduction via Surface Frustrated Lewis Pairs, ACS Catal., 2016, 6(10), 7109–7117,  DOI:10.1021/acscatal.6b01545 .
  129. L. L. Zeonjuk, N. Vankova, A. Mavrandonakis, T. Heine, G. Röschenthaler and J. Eicher, On the Mechanism of Hydrogen Activation by Frustrated Lewis Pairs, Chem. – Eur. J., 2013, 19(51), 17413–17424,  DOI:10.1002/chem.201302727 .
  130. M. G. Evans and M. Polanyi, Inertia and driving force of chemical reactions, Trans. Faraday Soc., 1938, 34, 11–24,  10.1039/TF9383400011 .
  131. J. N. Brønsted, Acid and Basic Catalysis, Chem. Rev., 1928, 5, 231–338,  DOI:10.1021/cr60019a001 .
  132. C. Jiang, O. Blacque, T. Fox and H. Berke, Reversible, Metal-Free Hydrogen activation by Frustrated Lewis Pairs, Dalton Trans., 2011, 40(5), 1091–1097,  10.1039/C0DT01255F .
  133. R. Dobrovetsky and D. W. Stephan, Catalytic Reduction of CO2 to CO by Using Zinc(II) and In Situ Generated Carbodiphosphoranes, Angew. Chem., Int. Ed., 2013, 52(9), 2516–2519,  DOI:10.1002/anie.201208817 .
  134. M. Wen, F. Huang, G. Lu and Z.-X. Wang, Density Functional Theory Mechanistic Study of the Reduction of CO2 to CH4 Catalyzed by an Ammonium Hydridoborate Ion Pair: CO2 Activation via Formation of a Formic Acid Entity, Inorg. Chem., 2013, 52(20), 12098–12107,  DOI:10.1021/ic401920b .
  135. P. M. Zimmerman, Z. Zhang and C. B. Musgrave, Simultaneous Two-Hydrogen Transfer as a Mechanism for Efficient CO2 Reduction, Inorg. Chem., 2010, 49(19), 8724–8728,  DOI:10.1021/ic100454z .
  136. B. Jiang, Q. Zhang and L. Dang, Theoretical Studies on Bridged Frustrated Lewis Pair (FLP) Mediated H2 Activation and CO2 Hydrogenation, Org. Chem. Front., 2018, 5(12), 1905–1915,  10.1039/C8QO00192H .
  137. M. Delarmelina, J. W. D. M. Carneiro, C. R. A. Catlow and M. Bühl, Design of CO2 Hydrogenation Catalysts Based on Phosphane/Borane Frustrated Lewis Pairs and Xanthene-Derived Scaffolds, Catal. Commun., 2022, 162, 106385,  DOI:10.1016/j.catcom.2021.106385 .
  138. M. Ghara, S. Pan and P. K. Chattaraj, A Theoretical Investigation on Boron–Ligand Cooperation to Activate Molecular Hydrogen by a Frustrated Lewis Pair and Subsequent Reduction of Carbon Dioxide, Phys. Chem. Chem. Phys., 2019, 21(38), 21267–21277,  10.1039/C9CP03756J .
  139. H. Sabet-Sarvestani, M. Izadyar and H. Eshghi, Molecular Electrostatic Potential at Nuclear Position as a New Concept in Evaluation of the Substitution Effects of Intramolecular B/N Frustrated Lewis Pairs in H 2 Splitting and CO2 Reduction, Int. J. Quantum Chem., 2020, 120(24), e26416,  DOI:10.1002/qua.26416 .
  140. M. Ghara and P. K. Chattaraj, A Computational Study on Hydrogenation of CO2 Catalyzed by a Bridged B/N Frustrated Lewis Pair, Struct. Chem., 2019, 30(3), 1067–1077,  DOI:10.1007/s11224-018-1264-4 .
  141. Z. Mo, E. L. Kolychev, A. Rit, J. Campos, H. Niu and S. Aldridge, Facile Reversibility by Design: Tuning Small Molecule Capture and Activation by Single Component Frustrated Lewis Pairs, J. Am. Chem. Soc., 2015, 137(38), 12227–12230,  DOI:10.1021/jacs.5b08614 .
  142. L. Liu, B. Lukose and B. Ensing, A Free Energy Landscape of CO2 Capture by Frustrated Lewis Pairs, ACS Catal., 2018, 8(4), 3376–3381,  DOI:10.1021/acscatal.7b04072 .
  143. M. Pu and T. Privalov, Ab Initio Molecular Dynamics with Explicit Solvent Reveals a Two-Step Pathway in the Frustrated Lewis Pair Reaction, Chem. – Eur. J., 2015, 21(49), 17708–17720,  DOI:10.1002/chem.201502926 .
  144. Z.-F. Zhang and M.-D. Su, Theoretical Study of the Activation Reaction of a Zr+/P-Based Frustrated Lewis Pair with Carbon Dioxide, J. Phys. Chem. A, 2022, 126(33), 5534–5544,  DOI:10.1021/acs.jpca.2c03602 .
  145. M. Ghara and P. K. Chattaraj, Can a Decrease in Anti-Aromaticity Increase the Dihydrogen Activation Ability of a Frustrated Phosphorous/Borane Lewis Pair: A DFT Study, Theor. Chem. Acc., 2020, 139(12), 183,  DOI:10.1007/s00214-020-02698-6 .
  146. J. H. Cavka; S. Jakobsen; U. Olsbye; N. Guillou; C. Lamberti; S. Bordiga and K. P. Lillerud A New Zirconium Inorganic Building Brick Forming Metal Organic Frameworks with Exceptional Stability.
  147. R. J. Marshall and R. S. Forgan, Postsynthetic Modification of Zirconium Metal-Organic Frameworks: Postsynthetic Modification of Zirconium Metal-Organic Frameworks, Eur. J. Inorg. Chem., 2016,(27), 4310–4331,  DOI:10.1002/ejic.201600394 .
  148. S. Ling and B. Slater, Dynamic Acidity in Defective UiO-66, Chem. Sci., 2016, 7, 4706–4712,  10.1039/C5SC04953A .
  149. K. Yang and J. Jiang, Computational Design of a Metal-Based Frustrated Lewis Pair on Defective UiO-66 for CO2 Hydrogenation to Methanol, J. Mater. Chem. A, 2020, 8(43), 22802–22815,  10.1039/D0TA07051C .
  150. E. Theuergarten, D. Schlüns, J. Grunenberg, C. G. Daniliuc, P. G. Jones and M. Tamm, Intramolecular Heterolytic Dihydrogen Cleavage by a Bifunctional Frustrated Pyrazolylborane Lewis Pair, Chem. Commun., 2010, 46(45), 8561,  10.1039/c0cc03474f .
  151. E. Theuergarten, J. Schlösser, D. Schlüns, M. Freytag, C. G. Daniliuc, P. G. Jones and M. Tamm, Fixation of Carbon Dioxide and Related Small Molecules by a Bifunctional Frustrated Pyrazolylborane Lewis Pair, Dalton Trans., 2012, 41(30), 9101,  10.1039/c2dt30448a .
  152. Y. Benseghir, A. Solé-Daura, D. R. Cairnie, A. L. Robinson, M. Duguet, P. Mialane, P. Gairola, M. Gomez-Mingot, M. Fontecave, D. Iovan, B. Bonnett, A. J. Morris, A. Dolbecq and C. Mellot-Draznieks, Unveiling the Mechanism of the Photocatalytic Reduction of CO2 to Formate Promoted by Porphyrinic Zr-Based Metal–Organic Frameworks, J. Mater. Chem. A, 2022, 10(35), 18103–18115,  10.1039/D2TA04164B .
  153. M. Faizan and R. Pawar, Alanine Boronic Acid Functionalized UiO-66 MOF as a Nanoreactor for the Conversion of CO2 into Formic Acid, J. Comput. Chem., 2023, 44(18), 1624–1633,  DOI:10.1002/jcc.27113 .
  154. M. P. Mitoraj, A. Michalak and T. Ziegler, A Combined Charge and Energy Decomposition Scheme for Bond Analysis, J. Chem. Theory Comput., 2009, 5(4), 962–975,  DOI:10.1021/ct800503d .
  155. R. Cauwenbergh, V. Goyal, R. Maiti, K. Natte and S. Das, Challenges and Recent Advancements in the Transformation of CO2 into Carboxylic Acids: Straightforward Assembly with Homogeneous 3d Metals, Chem. Soc. Rev., 2022, 51(22), 9371–9423,  10.1039/D1CS00921D .
  156. M. S. Ermolenko, S. Guillou and Y. L. Janin, Pyrazole-3/5-Carboxylic Acids from 3/5-Trifluoromethyl NH-Pyrazoles, Tetrahedron, 2013, 69(1), 257–263,  DOI:10.1016/j.tet.2012.10.034 .
  157. R. A. F. Tomás, J. C. M. Bordado and J. F. P. Gomes, P -Xylene Oxidation to Terephthalic Acid: A Literature Review Oriented toward Process Optimization and Development, Chem. Rev., 2013, 113(10), 7421–7469,  DOI:10.1021/cr300298j .
  158. T. V. Bukharkina, O. S. Grechishkina, N. G. Digurov and N. V. Krukovskaya, Kinetic Model of Ethyl Benzene Oxidation Catalysed by Manganese Salts, Org. Process Res. Dev., 2003, 7(2), 148–154,  DOI:10.1021/op9900986 .
  159. M. R. Mian, H. Chen, R. Cao, K. O. Kirlikovali, R. Q. Snurr, T. Islamoglu and O. K. Farha, Insights into Catalytic Hydrolysis of Organophosphonates at M–OH Sites of Azolate-Based Metal Organic Frameworks, J. Am. Chem. Soc., 2021, 143(26), 9893–9900,  DOI:10.1021/jacs.1c03901 .
  160. M. C. De Koning, M. Van Grol and T. Breijaert, Degradation of Paraoxon and the Chemical Warfare Agents VX, Tabun, and Soman by the Metal–Organic Frameworks UiO-66-NH2, MOF-808, NU-1000, and PCN-777, Inorg. Chem., 2017, 56(19), 11804–11809,  DOI:10.1021/acs.inorgchem.7b01809 .
  161. D. O. Wasik, A. Martín-Calvo, J. J. Gutiérrez-Sevillano, D. Dubbeldam, T. J. H. Vlugt and S. Calero, Enhancement of Formic Acid Production from Carbon Dioxide Hydrogenation Using Metal-Organic Frameworks: Monte Carlo Simulation Study, Chem. Eng. J., 2023, 467, 143432,  DOI:10.1016/j.cej.2023.143432 .

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3cs00267e
These authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2024