Vilakkathala U.
Krishnapriya
ab and
Cherumuttathu H.
Suresh
*bc
aChemical Science and Technology Division, CSIR-National Institute of Interdisciplinary Science and Technology, Thiruvananthapuram – 695019, Kerala, India
bResearch Centre, University of Kerala, Thiruvananthapuram, 695034, Kerala, India. E-mail: sureshch@gmail.com; sureshch@niist.res.in
cSrinivasa Ramanujan Institute for Basic Sciences, Kerala State Council for Science Technology and Environment, Kottayam, 686501, Kerala, India
First published on 4th November 2024
Activating atmospheric dinitrogen (N2), a molecule with a remarkably strong triple bond, remains a major challenge in chemistry. This theoretical study explores the potential of superbase phosphines, specifically those decorated with imidazolin-2-imine ((ImN)3P) and imidazolin-2-methylidene ((ImCH)3P) to facilitate N2 activation and subsequent hydrazine (H2NNH2) formation. Using density functional theory (DFT) at the M06L/6-311++G(d,p) level, we investigated the interactions between these phosphines and N2. Mono-phosphine–N2 complexes exhibit weak, noncovalent interactions (−0.6 to −7.1 kcal mol−1). Notably, two superbasic phosphines also form high-energy hypervalent complexes with N2, albeit at significantly higher energies. The superbasic nature and potential for the hypervalency of these phosphines lead to substantial N2 activation in bis-phosphine–N2 complexes, where N2 is “sandwiched” between two phosphine moieties through hypervalent P–N bonds. Among the phosphines studied, only (ImN)3P forms an exothermic sandwich complex with N2, stabilized by hydrogen bonding between the ImN substituents and the central N2 molecule. A two-step, exothermic hydrogen transfer pathway from (ImN)3P to N2 results in the formation of a bis-phosphine–diimine (HNNH) sandwich complex. Subsequent hydrogen transfer leads to the formation of a bis-phosphine–hydrazine (H2NNH2) complex, a process that, although endothermic, exhibits surmountable activation barriers. The relatively low energy requirements for this overall transformation suggest its potential feasibility under the optimized conditions. This theoretical exploration highlights the promise of superbase phosphines as a strategy for metal-free N2 activation, opening doors for the development of more efficient and sustainable nitrogen fixation and utilization methods.
The presence of vacant and filled d orbitals of appropriate energy and symmetry of transition metal complexes allows them to accept and back-donate the electron density with N2. This back-bonding weakens the highly stable N–N triple bond and is able to perform N2 activation. Main group elements, lacking such d orbitals, have traditionally been less effective in N2 activation. However, recent research has challenged this view, demonstrating the potential of metal-free catalysts for this crucial process. Extensive efforts have led to the development of various catalysts for N2 activation, including those based on transition metals (Fe, Mo, Au, etc.), main group metals (Bi, Si, etc.), and, more recently, metal-free systems.20–29 In the past three years, electrochemical ammonia production by main group metal-based catalysts has attracted greater attention in the field of N2 activation.6,30–33 While innovative approaches like electrochemical ammonia production and (photo)electrocatalysis have shown promise, progress has been hampered by reproducibility issues.34
Despite the predominance of metal-based catalysts, metal-free systems are gaining interest. In 2017, Bettinger et al. reported the photochemical extrusion of dinitrogen using phenylborylene under matrix isolation conditions.35 Braunschweig et al. reported that dinitrogen can easily be reduced by using borylene.36 Their work provides a pivotal moment in the scenario of dinitrogen activation using p-block elements and also offers that main group elements are also able to contribute in dinitrogen activation. Recently, p-block half-metal boron was found to be useful for the activation of N2.37 Among s-block elements, Li shows reactivity with N2 because of its extreme reduction potential.38,39 In 2021, Frenking et al. reported that low-valent Ca complexes can reduce dinitrogen.39 The exceptional durability and activity of metal-free catalysts, coupled with their economic and environmental benefits, underscore the importance of their development. Consequently, there is a growing emphasis on the design and synthesis of novel metal-free catalysts for N2 activation.
Very recently, Suresh and Krishnapriya40 showed the remarkable electron-donating properties of the moieties imidazolin-2-imine (ImN–), its methyl derivative (Im′N–), imidazolin-2-methylidene (ImCH–) and its methyl derivative (Im′CH–). These moieties substituted to benzene, pyridine, N-heterocyclic carbene and phosphine led to the design of extremely electron-rich ligands. Fareed and Suresh also showed that substitution of such moieties on the phenyl ring (Fig. 1b and c) can lead to more than a 4-fold increase in cation–π interaction energy compared to benzene.41 Dielmann et al. synthesized a variety of ImN-decorated extremely electron-rich phosphines.42 Such nonionic phosphorus(III) ligands showed superior electron donating ability to any other known phosphines in chemistry and they are described as a new class of superbases. Dielmann et al. investigated their unique electron-releasing character to describe their capability to capture and cleave CO2 molecules.43 With their superior electron-donating ability towards a core structure, these moieties are promising candidates in homogeneous catalysis, offering innovative possibilities for creating auxiliary ligands and advancing catalysis research.
Fig. 1 (a) C2 complex of the phosphine ligand synthesized by Leung et al. in 2021. (b) and (c) The cation–π receptors proposed by Fareed and Suresh in 2011. |
Recently, the diatomic molecule C2 has been structurally characterized by Leung et al. as a monoligated complex with the formula (ImaN)2(CH3)PC2 (Fig. 1a), where the N center of the imidazolidin-2-iminato groups (ImaN) is adorned with the bulky R group 2,6-diisopropylphenyl.44 Leung et al.'s discovery of a C2 complex of phosphine suggests an unprecedented reaction in chemistry and the driving force of this reaction could be attributed to the incredible electron-donating properties of the ImaN substituent which creates an extremely electron-rich phosphorus center in chemistry. The analysis of the electronic structure of (ImaN)2(CH3)PC2 with quantum chemical methods suggested the formation of the ligand-to-C2 (L → C2) coordination bond which induced a large charge migration from P towards C2. Despite the fact that ImCH- or Im'CH-based systems are yet to be synthesized, theoretical studies from our group suggest that phosphines decorated with these substituents are highly electron rich and they would behave as superbases (Fig. 2).
The focus of this study is on the metal-free N2 activation reactions facilitated by the superbases of (ImN)3P and (ImCH)3P types. The study will assess the electron richness of these ligands and investigate the formation of both L → N2 and L → N2 ← L complexes. We anticipate that the ability of phosphorus to achieve a hypervalent state will be key to the N2 activation process.
The molecular electrostatic potential (MESP) is a valuable tool in computational chemistry for visualizing and understanding the distribution of the electron density within a molecule.45–47 It provides insights into the regions of electrophilic and nucleophilic reactivities, intermolecular interactions, and the nature of chemical bonding.48–51 In this study, MESP topological analysis is employed to assess the electron richness of superbase ligands and to visualize changes in electron density distribution during the N2 activation process.
The MESP topological analysis was performed on the optimized geometries at the M06L/6-311++G(d,p) level of theory using the ‘cube’ and ‘potential’ keywords in Gaussian16. The MESP V(r) at any point with respect to the position vector r is defined as follows:
Fig. 3 Representation of the MESP isosurfaces of a set of phosphines at −20 kcal mol−1. Vmin values are given in kcal mol−1. Color code: blue, N; green, C; grey, H; rust brown, P. |
The introduction of imidazolin-2-imine (ImN–) and imidazolin-2-methylidene (ImCH–) moieties has been proposed as a strategy to design highly electron-rich substrates and ligands, leading to the classification of such ligands as superbases.40,60 Compared to PMe3, the superbases constructed through substitution with ImN–, Im′N–, ImCH–, and Im′CH– exhibit a notable increase in the magnitude of Vmin. For example, considering the Vmin values (Table 1), P(ImdN)3 emerges as the most electron-rich, demonstrating a 2.6-fold higher magnitude for Vmin compared to PH3, highlighting its exceptional electron richness. Previous studies have suggested that this new class of superbase phosphine ligands possesses an electron-donating ability comparable to that of N-heterocyclic carbenes.40
Phosphine | V min | Phosphine | V min |
---|---|---|---|
PH3 | −23.6 | P(ImdN)3 | −61.9 |
PMe3 | −39.0 | P(ImaCH)3 | −50.8 |
P(ImaN)3 | −47.6 | P(ImbCH)3 | −59.2 |
P(ImbN)3 | −52.7 | P(ImcCH)3 | −59.6 |
P(ImcN)3 | −56.8 | P(ImdCH)3 | −60.3 |
We further investigated complex formation involving a direct bond between the phosphorus of PR3 and the nitrogen of N2. Among all phosphines studied, only P(ImcN)3 and P(ImdN)3 showed the formation of such complexes with N2. These complexes, denoted as (ImcN)3P–N2 and (ImdN)3P–N2, are also depicted in Fig. 4. It is evident from the P–N distances that the P center forms bonds with five N centers, establishing the complex as a hypervalent phosphorus system. Notably, two of these P–N bonds are significantly shorter than those in free PR3, while another two exhibit nearly identical distances. Only one P–N bond is elongated by approximately 0.1 Å compared to free phosphine. The N–N distance of 1.28 Å clearly indicates substantial dinitrogen activation due to the hypervalent bonding with the phosphine ligand. However, the energetics suggest that both (ImcN)3P–N2 and (ImdN)3P–N2 are high-energy systems, lying 70.1 and 65.9 kcal mol−1, respectively, above free phosphine and N2. Therefore, their formation under ambient conditions is unlikely.
Fig. 5 Optimized structures of the dinitrogen bonded complexes of (a) PMe3 and (b) P(ImaN)3 ligands. Bond distances and N⋯HN hydrogen bond distances are shown in Å. |
Complex | ΔE | ΔG | D P–N1 | D P–N2 | D N–N |
---|---|---|---|---|---|
H3P–NN–PH3 | 87.4 | 111.2 | 1.59 | 1.59 | 1.46 |
Me3P–NN–PMe3 | 49.7 | 72.4 | 1.59 | 1.59 | 1.50 |
(ImaN)3P–NN–P(ImaN)3 | −13.3 | 18.5 | 1.63 | 1.63 | 1.48 |
(ImbN)3P–NN–P(ImbN)3 | −18.1 | 18.4 | 1.61 | 1.62 | 1.47 |
(ImcN)3P–NN–P(ImcN)3 | 13.5 | 51.0 | 1.61 | 1.61 | 1.47 |
(ImdN)3P–NN–P(ImdN)3 | 2.6 | 42.3 | 1.60 | 1.62 | 1.46 |
(ImaCH)3P–NN–P(ImaCH)3 | 5.1 | 35.2 | 1.65 | 1.65 | 1.49 |
(ImbCH)3P–NN–P(ImbCH)3 | 41.1 | 68.1 | 1.61 | 1.62 | 1.46 |
(ImcCH)3P–NN–P(ImcCH)3 | 22.1 | 53.1 | 1.64 | 1.64 | 1.43 |
(ImdCH)3P–NN–P(ImdCH)3 | 20.2 | 55.4 | 1.63 | 1.63 | 1.44 |
The energetics of the mono- and bis-phosphine complexes, summarized in Fig. 6, clearly demonstrate the superior N2 activation capability of phosphine superbases bearing ImN-type substituents on phosphorus compared to those with ImCH-type substituents. Notably, the formation of (ImaN)3P–NN–P(ImaN)3 and (ImbN)3P–NN–P(ImbN)3 is exothermic, releasing 13.3 and 18.1 kcal mol−1, respectively. This favorable energetic profile can be attributed to the additional stabilization provided by four N–H⋯N hydrogen bonds that surround the central N2 unit (Fig. 5b), a feature unique to these systems possessing N–H bonds.
Fig. 6 Relative energies of the mono phosphine and bis-phosphine complexes with N2. Energy data are given in kcal mol−1. |
In the bis-phosphine–N2 complex, dinitrogen is positioned between two phosphine moieties, forming two P–N bonds. This arrangement significantly alters the bonding environment at both the phosphorus and nitrogen centers. For instance, the three P–N bonds in P(ImaN)3 contract from 1.74, 1.71, and 1.72 Å to 1.69, 1.65, and 1.65 Å in the (ImaN)3P–NN–P(ImaN)3 complex, signifying stronger P–N bonds in the complex. Moreover, the two newly formed P–N bonds with N2 exhibit an even shorter bond length of 1.63 Å. These structural data unequivocally demonstrate that all five P–N bonds in the hypervalent bis-phosphine–N2 complex are substantially stronger than a typical P–N single bond. In essence, the remarkable hypervalent P–N bonding facilitates the activation of the N2 triple bond, effectively reducing it to an N–N single bond.
The P–N bonding in the bis-phosphine–N2 complex was further elucidated through NBO analysis, using Me3P–NN–PMe3 as a representative case. NBOs reveal that the N centers adopt an sp2 hybridized state (Fig. 7). Degenerate NBOs 1 and 2, depicted in Fig. 7, represent π-bonds between P and N, arising from the interaction of the lone pair on N (the ‘p’ orbital not involved in sp2 hybridization) with the Rydberg ‘d’ orbitals of the P center. This π-bonding interaction is predominantly contributed by the N center (89%), with the remaining contribution from the P center. Similarly, the degenerate NBOs 3 and 4 correspond to sp2 hybrid lone pair orbitals on each N centers. The NBO 5 describes the N–N σ-bond formed from the sp2 hybrid orbitals on the N centers. The degenerate NBOs 6 and 7 depict the σ-bonding orbitals between the third sp2 hybrid orbital of the N center and primarily the 3s orbital of the P center.
NBO analysis and the observed N–N bond lengths collectively confirm the reduction of the N–N bond to a single bond character in the bis-phosphine–N2 complexes. The calculated energy required for homolytic fission of Me3P–NN–PMe3 into two Me3P–N˙ radical fragments (66.1 kcal mol−1) and the analogous process for (ImaN)3P–NN–P(ImaN)3 (75.3 kcal mol−1) reveal a dramatic weakening of the N–N bond compared to its strength in dinitrogen (226 kcal mol−1). This substantial decrease in bond strength underscores the remarkable activation of N2 achieved through the formation of these hypervalent sandwich complexes with superbase phosphines.
The subsequent investigation, focusing on further N2 activation, will exclusively consider (ImaN)3P–NN–P(ImaN)3 due to its exothermic formation and the presence of stabilizing N–H⋯N hydrogen bonds.
Fig. 8 Optimized structures of the products and transition states for the formation of the diimine complex (ImaN′)3P–HNNH–P(ImaN′)3. P–N and N–N distances are shown in Å. |
The solvation energy-incorporated energy profile diagram for the hydrogen transfer within the (ImaN)3P–NN–P(ImaN) complex, leading to the activated diimine product (ImaN′)3P–HNNH–P(ImaN′)3, is presented in Fig. 9. This process involves two exothermic (and exergonic) hydrogen transfer steps. The initial hydrogen transfer, exothermic by 6.1 kcal mol−1, proceeds through ts1 with an activation energy of only 2.6 kcal mol−1. The resulting intermediate, (ImaN′)3P–NNH–P(ImaN)3, then undergoes a second hydrogen transfer via ts2. This step, with an activation barrier of 1.0 kcal mol−1, is even more exothermic, releasing 14.1 kcal mol−1. Overall, the reaction is exothermic by 20.2 kcal mol−1. When considering the free energy profile, both steps appear virtually barrierless, and the formation of the diimine complex (ImaN′)3P-NHNH-P(ImaN′)3 is exergonic by 18.0 kcal mol−1.
(ImaN′)3P–HNNH–P(ImaN′)3 → 2 (ImaN′)3P–HN˙ | (1) |
Fig. 9 Reaction energy profile diagram for hydrogen transfer occurs in (ImaN)3P–NN–P(ImaN)3 to form the diimine complex (ImaN′)3P–HNNH–P(ImaN′)3. |
The N–N bond strength of the diimine complex (ImaN′)3P–HNNH–P(ImaN′)3 was evaluated through the energetics of reaction (1). Cleavage of the N–N bond, yielding the radical (ImaN′)3P–HN˙, requires a dissociation energy of 45.6 kcal mol−1, highlighting the significant activation of the N–N bond. Notably, the N–N bond in the diimine is weakened by 29.7 kcal mol−1 compared to that observed in the dinitrogen complex (ImaN)3P–NN–P(ImaN)3.
Fig. 10 Optimized structures of the products and transition states for the formation of the hydrazine complex (ImaN′′)3P–H2NNH2–P(ImaN′′)3. P–N and N–N distances are shown in Å. |
The solvation-corrected energy and free energy profiles (Fig. 11) reveal that this reaction has an activation energy of 8.6 kcal mol−1 and an activation free energy of 6.7 kcal mol−1. It is also endothermic by 7.0 kcal mol−1 and endergonic by 9.8 kcal mol−1.
Fig. 11 Reaction energy profile diagram for hydrogen transfer occurs in (ImaN′)3P–HNNH–P(ImaN′)3 to form the hydrazine complex (ImaN′′)3P–H2NNH2–P(ImaN′′)3. |
A subsequent hydrogen transfer within the intermediate (ImaN′′)3P–HNNH2–P(ImaN′)3 leads to the hydrazine complex (ImaN′′)3P–H2NNH2–P(ImaN′′)3via ts4. This step exhibits activation energies of 11.3 kcal mol−1 (energy) and 6.6 kcal mol−1 (free energy). Overall, the formation of the hydrazine complex from the diimine complex is endothermic by 15.8 kcal mol−1 and endergonic by 16.1 kcal mol−1. However, both the activation energy and activation free energy are readily surmountable under ambient conditions.
We also modelled the dissociation of both (ImaN′′)3P ligands from the hydrazine complex (ImaN′′)3P–H2NNH2–P(ImaN′′)3 to generate the hydrazine H2NNH2. Since two P–N bonds are cleaved in this dissociation, a stepwise dissociation is considered. The first P–N bond dissociation (reaction (2)) occurs with an energy change of 30.9 kcal mol−1 and a free energy change of 15.2 kcal mol−1. The second P–N bond dissociation, releasing hydrazine (reaction (3)), occurs from (ImaN′′)3P–H2NNH2 and has an energy change of 28.7 kcal mol−1 and a free energy change of 12.4 kcal mol−1.
(2) |
(ImaN′′)3P–H2NNH2 → (ImaN′′)3P + H2NNH2 | (3) |
Comparing the energy profiles for diimine formation (Fig. 9) and hydrazine formation (Fig. 11) reveals that the diimine complex is both a kinetically and thermodynamically favoured product. However, the transformation from the diimine complex to the hydrazine complex, while endothermic, remains feasible due to its relatively low barrier of 18.3 kcal mol−1. Furthermore, by strategically manipulating reaction conditions such as temperature, pressure, and reactant/product concentrations in accordance with Le Châtelier's principle, the equilibrium can be shifted to favour the formation of hydrazine, despite its thermodynamically less favoured nature.
DFT calculations at the M06L/6-311++G(d,p) level reveal that bis-phosphine complexes, forming hypervalent sandwich structures with N2, exhibit the most promising N2 activation. This activation is driven by the formation of strong P–N bonds with double bond character and a concomitant weakening of the N–N bond, as evidenced by structural and bonding analyses. The superbasic nature of the ImN-type substituents on the phosphine ligands plays a crucial role in enhancing their affinity for N2 and stabilizing the resulting complexes.
We propose a two-step hydrogen transfer pathway for the (ImN)3P–N2–P(ImN)3 sandwich complex, leading to the exothermic formation of a diimine intermediate with a significantly weakened N–N bond. Further hydrogen transfer results in the formation of a hydrazine complex, a process that, while endothermic, exhibits surmountable activation barriers.
This research underscores the potential of ImN-type superbase phosphine ligands in facilitating nitrogen activation and transformation. The ability to controllably convert dinitrogen into more reactive intermediates like diimine and hydrazine opens new avenues for the design of metal-free catalytic systems for nitrogen fixation and other applications in chemical synthesis. Experimental validation of these computational findings is the next crucial step in realizing the potential of this novel approach.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4dt02703e |
This journal is © The Royal Society of Chemistry 2024 |