DOI:
10.1039/D5QI02041G
(Research Article)
Inorg. Chem. Front., 2026, Advance Article
Catalytic activation of nitrous oxide: boryl versus hydride nickel complexes
Received
7th October 2025
, Accepted 27th November 2025
First published on 28th November 2025
Abstract
The selective reduction of nitrous oxide (N2O), a potent greenhouse gas with harmful effects on the ozone layer, remains a significant challenge in small-molecule activation. Herein, we report the efficient deoxygenation of N2O using bis(phosphino)boryl-nickel hydride (3) and bis-boryl nickel complexes (4 and 5) under mild conditions (1–2 bar N2O, 2–5 mol% catalyst loading, and 25 °C). Both catalytic systems exhibit high activity in the presence of boranes and diboranes, achieving complete N2O conversion within 30 minutes using catecholborane as the reductant. Mechanistic investigations, including stoichiometric experiments, kinetic studies, and density functional theory (DFT) calculations support the formation of nickel boroxide intermediates, (RPBP)Ni–OBR2, as key species within the catalytic cycle, while pathways involving nickel hydroxide species, (RPBP)Ni–OH, are disfavored. These results provide valuable mechanistic insights into key aspects of N2O reduction chemistry thereby enabling the rational design of transition-metal catalysts for the activation of small-molecules.
Introduction
Atmospheric levels of greenhouse gases have steadily increased since preindustrial time due to human activities, with carbon dioxide, methane, and nitrous oxide being the primary contributors to the climate change.1 According to the Intergovernmental Panel on Climate Change (IPCC), the concentrations of methane and nitrous oxide have reached unprecedented levels while current concentrations of carbon dioxide are higher than any point in history.1,2 Commonly known as laughing gas, nitrous oxide is often associated with recreational use; however, N2O is a potent greenhouse gas, with a global warming potential 300 times greater that CO2 and it is also one of the main contributors to the depletion of the ozone layer.3 Nitrous oxide naturally occurs in the atmosphere as part of the Earth's nitrogen cycle, and has numerous natural sources. However, approximately 40% of total nitrous oxide emissions are linked to human sources including agricultural practices, fuel combustion and the industrial production of nitric and adipic acid.4 In this context, strategies to control N2O emissions have spurred numerous investigations in recent years. N2O typically reacts as an oxidizing agent, transferring oxygen and releasing N2, although, an alternative reactivity, serving as a diazo transfer reagent, has also been reported.5,6 However, due to its inert nature, harsh reaction conditions such as elevated temperature or pressure, in the presence of heterogeneous catalysts, are usually required to achieve efficient conversion.7 Frustrated Lewis pairs (FLPs) offer pathways for the activation of N2O under mild conditions through the formation of oxo compounds.8 Efficient catalytic deoxygenation of N2O has been achieved using disilanes as reducing reagents in combination with catalytic amounts of fluorides or alkoxides, as demonstrated by Cantat and coworkers.9 The Cornella group has shown that low-valent Bi(I) species can effectively reduce N2O using pinacolborane as the reducing agent.10 Recently, Paradis and coworkers accomplished the catalytic reduction of N2O using metal-free P(III)/P(V)
O catalysis with phenyl silane as reductant.11 In Nature, nitrous oxide is converted into dinitrogen and water by nitrous oxide reductase (NOR), which utilizes a Cu4S cluster to carry out this reaction during the denitrification process.12 Drawing inspiration from this, many research groups have made efforts to design transition metal systems that emulate this natural process.13–26 Their reactivity is typically associated with oxygen atom transfer reactions to form M
O bonds or metal–hydroxide, alkoxide or aryloxide species by insertion into M–H or M–C bonds, with release of N2. Milstein and co-workers reported the efficient catalytic hydrogenation of N2O mediated by a Ru–H complex, where the O-insertion into the Ru–hydride bond was identified as the key step.14,15 Similarly, the mechanism proposed by Cornella and his team for the synthesis of alcohols using N2O, catalyzed by nickel, suggests an O-insertion into alkyl and aryl nickel species.16,17 Suárez and co-workers reported the catalytic N2O reduction with pinacolborane using a CNP–Ni hydride complex, DFT analysis suggested the formation of NiOH species, resulting from the insertion of N2O into the Ni–H bond.18–20 Lee, Baik and coworkers reported the stepwise reduction sequence from nitrate to N2, using a (PNP)Ni complex.21 Interestingly, the Mo group has presented a new perspective by proposing a cooperative iron–silicon mechanism in the deoxygenation of N2O, facilitated by an iron complex containing a bis(silylene) amido ligand.22 Recently, Chaplin and co-workers reported the catalytic reduction of N2O by insertion into a Cu–B bond, leading to the boroxide derivative, detected by 11B NMR during the catalytic reaction.23 Encouraged by these results, we envisioned as a stimulating approach to explore the catalytic reduction of N2O by O-transfer into the Ni–H, Ni–C, and Ni–B bonds of bis(phophino)boryl nickel complexes (RPBP)NiX (R = tBu, Cy; X = CH3, H, Bcat (cat = catecholato), Bcat′ (cat′ = 4-methylcatecholato) (2–5)) developed in our group (Fig. 1).27,28 Herein, we present a comparative analysis of the catalytic reduction of N2O using these nickel species, alongside stoichiometric reactivity studies with various boranes and diborane compounds, kinetic analyses, and DFT studies. These results provide insights into the active species involved in the reaction and propose a plausible mechanism for the activation of N2O.
 |
| | Fig. 1 Top: Kinetic plots of ln[HBpin] vs. initial time (s) in C6D6 using catalyst 2a (5 mol% catalyst loading); PN2O = 4 bar; [HBpin] = 0.16 mM; T = 288, 293, 298, 303 and 308 K. Bottom: Eyring plot of ln(k/T) vs. 1/T, from 288 to 308 K, ΔH‡ = 15.7 ± 0.7 kcal mol−1, ΔS‡ = −9.4 ± 0.6 cal mol−1 K−1, ΔG‡ = 18.5 ± 0.9 kcal mol−1 at 25 °C. | |
Results and discussion
The synthesis of complexes (tBuPBP)NiX (X = Br (1a), CH3 (2a), H (3), Bcat (4a)) was previously reported by us27,28 (Scheme 1). Compound (CyPBP)NiCH3 (2b) was synthesized by reaction of (CyPBP)NiBr (1b) complex with MgMe2(tmeda) (tmeda = N,N,N′,N′-tetramethylethane-1,2-diamine) in pentane (see SI for details). In this reaction, MgBr2(tmeda) is formed alongside 2b, and its low solubility in pentane prevents the regeneration of 1b, which presents a challenge for the synthesis of these species (Scheme 1).29 Although we have previously synthesized hydride complex 3, all attempts to prepare the hydride complex with Cy groups on phosphorus were unsuccessful. Bis-boryl nickel complexes 4b and 5a,b were prepared, using a procedure similar to that described for 4a, by treatment of the corresponding methyl complex (RPBP)NiCH3 (R = tBu (2a), Cy (2b)) with diboranes B2cat′2 and B2cat2 (Scheme 1). Compound 5a was isolated in moderate yield, and its B{1H} spectrum displayed the characteristic Ni-boryl resonances at 50 and 60 ppm. The 31P{1H} NMR spectrum exhibited a signal at 117 ppm, closely matching that observed for 4a.28 Complexes 4b and 5b were fully characterized by NMR; however, all attempts to isolate them were unsuccessful due to their high sensitivity to moisture and air (see SI).
 |
| | Scheme 1 Nickel complexes tested in the reactivity with N2O. | |
Upon treatment with N2O, we observed a distinct reactivity depending on the nickel species. While the reaction of methyl complexes 2a and 2b with N2O did not yield any insertion product, even after heating at 70 °C for 12 hours, the nickel hydride complex tBuPBPNiH (3) decomposed in the presence of N2O (see SI). In contrast, the bis(boryl) nickel derivative 4a reacts cleanly to form the nickel boroxide complex tBuPBPNiOBcat (6a) (Scheme 1; see SI for details). Complete transformation is confirmed by the disappearance of the signals corresponding to 4a in the 31P{1H} at δ 117 ppm along with the appearance of a new resonance at δ 84 ppm for the new species 6a. The 11B{1H} spectrum exhibits a resonance upfield at δ 24 ppm expected for the boroxide group. Similar reactivity was observed with analogous species 5a, however, for complexes 4b and 5b, the reaction was not clean leading to the formation of several products (Scheme 1; see SI for details). To evaluate the feasibility of a hypothetical catalytic cycle, B2Cat2 was added to 6a resulting in the recovery of bis(boryl) nickel species 4a and the release of O(Bcat)2 (Scheme 2; see SI for details). Based on these promising results, 4a (generated in situ from 2a and B2Cat2)28 was initially selected to evaluate the catalytic reduction of N2O with B2Cat2 (Table 1).
 |
| | Scheme 2 O-insertion of N2O into a Ni–B bond of 4a and further reactivity with B2Cat2, represented as a hypothetical catalytic cycle. | |
Table 1 Summary of the catalytic reduction of N2O to N2
a
| Entry |
[Ni] |
Solvent |
T |
Reductant |
Cat. loading mol% b |
Time (h) |
Conv. d (%) |
| Reaction conditions, unless otherwise specified: 2 bar N2O, C6D6 (400μL), [HBcat] 0.2 mM; [HBpin] 0.2 mM; [B2Cat2] = 0.2 mM. The catalyst loading is calculated based on the amount of catalyst related to the amount of reductant used. Reaction carried out in a Fisher Porter reactor; PN2O = 1 bar. Conversion was determined by 11B{1H} NMR spectroscopy using BH3(NMe3) as an internal standard. The formation of N2 and H2 was detected by GC-MS analysis of the headspace gas and 1H NMR spectroscopy, respectively (see SI). |
| 1 |
2a |
C6D6 |
25 |
B2Cat2 |
5 |
19 |
76 |
| 2 |
2a |
Toluene |
25 |
B2Cat2 |
5 |
19 |
75 |
| 3 |
2a |
C6H5F |
25 |
B2Cat2 |
5 |
19 |
75 |
| 4 |
2a |
THF |
25 |
B2Cat2 |
5 |
10 |
— |
| 5 |
2a |
CH2Cl2 |
25 |
B2Cat2 |
5 |
10 |
— |
| 6 |
2a |
C6D6 |
50 |
B2Cat2 |
5 |
10 |
99 |
| 7 |
2a |
C6D6 |
50 |
B2Cat′2 |
5 |
6 |
99 |
| 8 |
2a |
C6D6 |
50 |
B2pin2 |
5 |
10 |
0 |
| 9 |
2b |
C6D6 |
50 |
B2Cat2 |
5 |
4 |
99 |
| 10 |
2b |
C6D6 |
50 |
B2Cat′2 |
5 |
3.5 |
99 |
| 11 |
2a |
C6D6 |
25 |
HBcat |
5 |
0.5 |
99 |
| 12c |
2a |
C6D6 |
25 |
HBcat |
2 |
1 |
99 |
| 13 |
2a |
C6D6 |
25 |
HBpin |
5 |
2 |
99 |
| 14 |
2a |
C6D6 |
25 |
9-BBN |
5 |
2 |
0 |
| 15 |
2b |
C6D6 |
25 |
HBcat |
5 |
19 |
99 |
| 16 |
3 |
C6D6 |
25 |
HBcat |
5 |
0.5 |
99 |
| 17 |
3 |
C6D6 |
25 |
HBpin |
5 |
1.5 |
99 |
After 19 hours at 25 °C with 5 mol% of precatalyst 2a and 2 bar of N2O, 76% conversion of N2O into N2 and O(BCat)2 was observed (Table 1, entry 1). Reactions carried out in toluene or fluorobenzene led to similar conversions (Table 1, entries 2 and 3), while THF and CH2Cl2 showed a detrimental effect (Table 1, entries 4 and 5).30 When the reaction was conducted at 50 °C, in C6D6, full conversion was achieved after 10 h (Table 1, entry 6). The use of a slightly more soluble diborane B2(4-methylcatecholato)2 (B2Cat2′) resulted in a faster reaction (Table 1, entry 7), while no reaction was observed with bis(pinacolato)diborane (B2pin2), most likely due to steric hindrance between the pinacol groups on boron and the tert-butyl substituents on phosphorous (Table 1, entry 8). Using complex CyPBPNiCH3 (2b) as precatalyst, reduced the reaction time to just 4 hours or 3.5 h, depending on the diborane used (Table 1, entries 9 and 10). Notably, when catecholborane (HBcat) was used as the reductant, the reaction reached 99% conversion after just 30 minutes at 25 °C with 2a as precatalyst (Table 1, entry 11). An identical conversion was also achieved within just 1 hour using only 2 mol% catalyst loading in a Fischer Porter reactor (Table 1, entry 12). Similarly, full conversion is obtained after 1.5 hours at 25 °C using pinacolborane (HBpin) (Table 1, entry 13); however, no reaction occurred with 9-BBN as reductant (Table 1, entry 14). Precatalyst 2b proved to be less efficient with HBcat, reaching 99% conversion after 19 h (Table 1, entry 15). We tentatively attribute this low performance to the reduced stability of the nickel hydride intermediate that might be involved in this reaction (further discussion can be found below).
Notably, catalyst 3 exhibited similar conversions to 2a, using HBcat and HBpin as reductants (Table 1, entries 16 and 17), suggesting that analogous catalytic species may be operating under these conditions. Control experiments confirmed that no reaction occurs in the absence of the nickel catalyst. These results indicate that both nickel boryl (4a) and hydride (3) complexes are highly effective catalysts for the N2O deoxygenation reaction.
To gain a deeper understanding on the nature of the nickel species involved in these catalytic reactions, individual stoichiometric reactions potentially occurring in the catalytic cycle were investigated using complexes 2–5. As shown before, the reaction of 6a with B2Cat2 results in the formation of 4a accompanied by the generation of O(Bcat)2 (Scheme 3). A similar reactivity is observed when HBcat is used; however, in this case, the nickel hydride derivative 3 is formed (Scheme 3).
 |
| | Scheme 3 Reactivity of tBuPBPNiOBcat (6a) with B2Cat2 to form 4a (right) and with HBcat to form 3 (left). | |
Treatment of 3 with N2O results in a mixture of unidentified products (pathway a in Scheme 4, see SI). The formation of the product from N2O insertion into the Ni–H bond was not observed, even when the reaction was monitored by 1H and 31P NMR spectroscopy at variable temperature (from −60 °C to 25 °C). In contrast, when 3 reacts with N2O in the presence of HBcat, the reaction cleanly produces complex 6a, along with release of H2 and N2 (pathway b in Scheme 4). We hypothesized that the reaction begins with the formation of complex tBuPBPNiH2Bcat (8) from nickel hydride 3 and HBcat (pathway c in Scheme 4), which then reacts with N2O to ultimately produce 6a, accompanied by the elimination of H2 and N2 (pathway d in Scheme 4). To prove our hypothesis, 8 was prepared, by reacting 3 with HBcat (or alternatively by reaction of 4a with H2),22 then, N2O was added. The generation of 6a and release of H2 was confirmed by 1H, 31P{1H} and 11B{1H} NMR spectroscopy (see SI).31 Based on these experimental results we propose that nickel boroxide complex 6a may serve as a key intermediate in the catalytic reaction and, the potential pathway involving the initial generation of nickel hydroxide species can be ruled out.
 |
| | Scheme 4 Reactivity studies of 3 with N2O and HBcat. | |
Following a detailed analysis of each step in the catalytic reaction, we carried out kinetic studies to assess the dependence of the reaction rate on each component involved. The experiment was performed using precatalyst 2a, 2 bar of N2O and HBpin (0.08 mmol) in C6D6. The time-dependence of the conversion of HBpin into O(Bpin)2 was monitored by 11B{1H} NMR spectroscopy. Initial experiments revealed that the reaction is first-order in HBpin and zero order with respect to N2O. Kinetic studies varying the concentration of 2a showed a first-order dependence on the catalyst consisting with an overall rate law of kobs[2a][HBpin] (see SI for details). To estimate experimentally the thermodynamic parameters of the overall energy barrier, (ΔH‡, ΔS‡, and ΔG‡), the reaction was monitored at various temperatures ranging from 15 to 35 °C (Fig. 1).
An Eyring analysis, based on a plot of ln(k/T) versus 1/T, yielded an activation enthalpy (ΔH‡) of 15.7 ± 0.7 kcal mol−1 and activation entropy (ΔS‡) of −9.4 ± 0.6 cal mol−1 K−1. The corresponding Gibbs free energy of activation (ΔG‡) at 298 K was calculated to be 18.5 ± 0.9 kcal mol−1.
The reduction of N2O mediated by PBP-supported Ni complexes was also explored using DFT methods. Initial investigations involved the reaction between hydride species 3 and N2O, where nucleophilic attack of the hydride moiety to the terminal nitrogen on N2O starts the catalytic transformation.32 This step requires 17.7 kcal mol−1 (TS1, Fig. 2), which is a considerably lower value than that calculated for recent examples involving picoline-derived CNP ligands,18 probably due to the increased hydride character of 3 as a consequence of the strong trans influence boryl fragment.29,33 Following the hydride transfer, intermediates Int1 and Int1′ (−12.0 and −14.7 kcal mol−1, respectively) were found in the potential energy surface, and represent a formal insertion of N2O into the Ni–H bond. From these complexes, attempts to extrude N2 with concomitant formation of a hypothetical and experimentally unobserved Ni hydroxo species were pursued by a number of methods. Some of these include intramolecular 4-membered transition states (similar to what was previously found for related species),18 participation of the PBP ligand, hydride transfer to the central nitrogen atom on N2O, or processes containing two N2O molecules (more details can be found in the SI). However, all these pathways led to very high energy kinetic barriers (≥30 kcal mol−1), in line with the experimental observations and the lack of a stable isolable or detectable Ni–OH complex. Therefore, the inclusion of a hydroborane to continue with the N2O reduction process is necessary, as experimentally found. HBpin was considered for these calculations, in order to compare the calculated energy barrier with that derived from the kinetic experiments. Similar to that described above for N2O, the hydride ligand on 3 can attack the Lewis acidic boron atom of HBpin via TS-HBpin (9.0 kcal mol−1), but the resulting hydridoborate 9 is unstable towards dissociation (3.8 kcal mol−1), leaving HBpin available in the reaction mixture (nonetheless, N2O activation mediated by 9 has also been computationally explored, see SI for more information). Thus, the oxygen atom in intermediate Int1′ can attack HBpin via TS2, 0.4 kcal mol−1 above the energy reference. The result of this interaction is the corresponding adduct Int2 (−2.9 kcal mol−1), where the HBpin molecule and the O–N
N–H moiety establish a 6-membered ring stabilized by an O⋯H hydrogen bond (1.87 Å). Rotation of the HBpin molecule places the hydridic H atom close to the proton of the O–N
N–H group, which leads to the formation of dihydrogen with concomitant liberation of N2 via TS3. This transition state is the rate-determining step in the proposed mechanism with an energy barrier of 19.4 kcal mol−1, in very good agreement with that obtained experimentally (18.5 ± 0.9 kcal mol−1). The preorganization imposed by the hydrogen bond in Int2 leads to a rigid transition state where the orientation of the HBpin and ONNH fragments are key for the formation of the H2 molecule. Expectedly, the entropy for the rate-determining step of this mechanism is negative (−3.9 cal mol−1 K−1), in accordance with the experimental value (−9.4 ± 0.6 cal mol−1 K−1). The formation of thermodynamically stable by-products translates into an abrupt decrease in energy after liberation of N2 and H2, placing boroxide complex NiOBpin 101.6 kcal mol−1 below the energy reference. Regeneration of catalytically active species 3 derives from the approach of another equivalent of HBpin. In a similar manner to what was described for TS2, the oxygen atom in NiOBpin can attack the boron atom of HBpin via TS4 (−85.9 kcal mol−1). The participation of two Bpin fragments and the associative character of this transition state involves a negative entropic term (−22.9 cal mol−1 K−1). However, instead of obtaining the expected adduct where oxygen remains bound to the metal center, a subtle shift takes place to yield Int3, which can be described as a σ-borane complex, with Ni–H and B–H bond distances of 1.65 and 1.35 Å, respectively. From this geometry, decoordination of the bis(boryl)ether only requires 3.5 kcal mol−1 through TS5, which gives back hydride species 3.
 |
| | Fig. 2 Computed Gibbs energy profile in benzene (SMD) for the reduction of N2O mediated by complex 3 and HBpin. Relative Gibbs energies computed at 298 K and 1 M are given in kcal mol−1. The Gibbs energy of 3 + N2O + HBpin (2 equiv.) has been taken as zero energy. All data have been computed at the SMD-PBE0-D3/def2-TZVP/def2-QZVP&SDD//SMD-PBE0-D3/6-31G(d,p)&SDD(+f) level. | |
Conclusion
In summary, we have synthetized a number of nickel methyl, hydride and boryl complexes supported by diphosphino boryl ligands to test their reactivity with nitrous oxide. We have demonstrated that R–PBP–nickel boryl (4–5) and hydride (3) complexes are active catalysts for the deoxygenation of N2O with release of N2, using boranes and diboranes as reductants, under mild conditions. Stoichiometric experiments revealed that oxygen transfer from N2O to the Ni–B bond affords novel nickel boroxide species, PBPNi–OBR2 (6–7), whereas reaction with the nickel hydride derivative leads to unidentified decomposition products. Furthermore, our investigations establish that in both cases nickel boroxides serve as key intermediates within the catalytic cycle, thereby excluding the participation of nickel hydroxide species when nickel hydride species 3 is used. Detailed kinetic analyses combined with DFT calculations indicate that the catalytic cycle is initiated by the insertion of N2O into the nickel–hydride bond, generating a Ni–ONNH intermediate. Rather than forming nickel–hydroxide species and liberating N2, this intermediate reacts directly with HBcat, releasing N2 and H2, and generating Ni–OBcat, which serves as the catalytically active species. Overall, this work provides a deeper understanding of N2O deoxygenation pathways and offer valuable information for the design of versatile platforms for small molecules activation reactions.
Methods
General procedure for the catalytic reduction of N2O with RPBPNiX species 2a (R = tBu; X = CH3), 2b (R = Cy; X = CH3) and 3 (R = tBu; X = H)
To a high-pressure J. Young valve NMR tube containing the catalyst or precatalyst (5 or 2 mol%), C6D6 (500 μL) and borane or diborane (0.08 mmol) were added and the J. Young valve NMR tube was degassed via three freeze–pump–thaw cycles. Then, the J. Young valve NMR tube was backfilled with N2O gas (2 bar). The reaction mixture was placed in a water bath at 25 °C and monitored by 11B{1H} NMR spectroscopy using BH3(NMe3) as an internal standard. All reactions were performed in duplicate in deuterated solvent.
Computational details
Calculations were performed using the PBE0 functional,34 as implemented in Gaussian 0935 along with Grimme's D3 dispersion correction.36 Geometry optimizations were performed in solution (solvent = benzene, ε = 2.27) using the continuum SMD model37 and basis set 1 (BS1). BS1 uses the double-ζ 6-31G(d,p)38 basis set for the H, C, N, O, B, and P atoms and the scalar relativistic Stuttgart–Dresden SDD pseudopotential39 and its associated double-ζ basis set, complemented with a set of polarization functions, for the Ni atom.40 The nature of the stationary points was confirmed by frequency analysis. Connections between the transition states and the minima were checked by perturbing the transition state geometry along the TS coordinate and optimizing until the corresponding minima. All energies in solution were corrected by single-point calculations with the larger basis set 2 (BS2) including triple-ζ def2TZVP basis set for the H, C, N, O, B, and P atoms, and def2QZVP for the Ni atom.41 The scalar relativistic Stuttgart–Dresden SDD pseudopotential and its associated basis set for Ni was used in the energy profile calculations. Gibbs energies in benzene were calculated at 298.15 K. Gibbs energy corrections were obtained based on vibrational frequencies of the BS1 optimized structures using the quasi-harmonic approximation. Thermal contributions to the Gibbs energies were corrected by employing the approximation described by Grimme, where entropic terms for frequencies below a cut-off of 100 cm−1 were calculated using the free-rotor approximation.42 The GoodVibes program developed by Paton and Funes-Ardoiz was employed to introduce these corrections.43 All reported energies in the main text correspond to PBE0-D3/BS2 Gibbs energies in benzene solvent (1 M) at 298.15 K in kcal mol−1. Structure visualization was performed with Chemcraft software.44
Author contributions
C. J. L. G., E. M. F. and A. R. performed the experimental work. R. P. performed the analysis of the kinetic data. P. R. performed the DFT calculations. R. P., P. R. and A. R. wrote the manuscript. A. R. conceived the project and directed the work. All authors have given approval to the final version of the manuscript.
Conflicts of interest
There are no conflicts to declare.
Data availability
The data supporting this article have been included as part of the supplementary information (SI). Supplementary information: synthesis and characterization details; experimental procedures, kinetic studies, compound characterization data, NMR spectra (PDF) and extended computational details (PDF). Cartesian coordinates and energies of optimized structures (XYZ). See DOI: https://doi.org/10.1039/d5qi02041g.
Acknowledgements
This work was supported by the Spanish MCIN/AEI/10.13039/501100011033 (PID2022-141925NB-I00 and RED2022-134287-T). C. J. Laglera-Gándara thanks for a Margarita Salas grant financed by the European Union-Next Generation EU, Ministry of Universities and Recovery, Transformation and Resilience Plan, through a call from University of Oviedo (Grant MU-21-UP2021-030 53307942). The use of the computational facilities of the Supercomputing Center of Galicia (CESGA) is gratefully acknowledged.
References
- IPCC, in Climate Change 2023: Synthesis Report. Contribution of Working Groups I, II and III to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change, ed. Core Writing Team, H. Lee and J. Romero, IPCC, Geneva, Switzerland, 2023, 184 pp., DOI:10.59327/IPCC/AR6-9789291691647.
- J. Hansen and M. Sato, Greenhouse Gas Growth Rates, Proc. Natl. Acad. Sci. U. S. A., 2004, 101, 16109–16114 CrossRef CAS.
- A. R. Ravishankara, J. S. Daniel and R. W. Portmann, Nitrous Oxide (N2O): The Dominant Ozone-Depleting Substance Emitted in the 21st Century, Science, 2009, 326, 123–125 CrossRef CAS PubMed.
- H. Tian, R. Xu, J. G. Canadell, R. L. Thompson, W. Winiwarter, P. Suntharalingam, E. A. Davidson, P. Ciais, R. B. Jackson and G. Janssens-Maenhout, et al., A Comprehensive Quantification of Global Nitrous Oxide Sources and Sinks, Nature, 2020, 586, 248–256 CrossRef CAS PubMed.
- K. Severin, Synthetic Chemistry with Nitrous Oxide, Chem. Soc. Rev., 2015, 44, 6375–6386 RSC.
- A. Genoux and K. Severin, Nitrous Oxide as Diazo Transfer Reagent, Chem. Sci., 2024, 15, 13605–13617 RSC.
- V. N. Parmon, G. I. Panov and A. S. Noskov, Nitrous Oxide in Oxidation Chemistry and Catalysis: Application and Production, Catal. Today, 2005, 100, 115–131 CrossRef CAS . A number of metal (Cu, Ru, …) surfaces catalyze this transformation at cryogenic temperatures and (sub)ambient pressures. See: W. A. Brown, R. K. Sharma, D. A. King and S. Haq, Adsorption and Reactivity of NO and N2O on Cu{110}: Combined RAIRS and Molecular Beam Studies, J. Phys. Chem., 1996, 100, 12559–12568 CrossRef.
- D. W. Stephan and G. Erker, Frustrated Lewis Pair Chemistry of Carbon, Nitrogen and Sulfur Oxides, Chem. Sci., 2014, 5, 2625–2641 RSC.
- L. Anthore-Dalion, E. Nicolas and T. Cantat, Catalytic Metal-Free Deoxygenation of Nitrous Oxide with Disilanes, ACS Catal., 2019, 9, 11563–11567 CrossRef CAS.
- Y. Pang, M. Leutzsch, N. Nöthling and J. Cornella, Catalytic Activation of N2O at a Low-Valent Bismuth Redox Platform, J. Am. Chem. Soc., 2020, 142, 19473–19479 CrossRef CAS PubMed.
- R. Zhou, V. Medvarić, T. Werner and J. Paradies, Metal-Free Reduction of Nitrous Oxide via PIII/PV
O Cycling: Mechanistic Insights and Catalytic Performance, J. Am. Chem. Soc., 2025, 147, 37879–37884 CrossRef CAS PubMed. - A. Pomowski, W. Zumft, P. M. H. Kroneck and O. Einsle, N2O Binding at a [4Cu:2S] Copper–Sulphur Cluster in Nitrous Oxide Reductase, Nature, 2011, 477, 234–237 CrossRef CAS PubMed.
- W. B. Tolman, Binding and Activation of N2O at Transition-Metal Centers: Recent Mechanistic Insights, Angew. Chem., Int. Ed., 2010, 49, 1018–1024 CrossRef CAS PubMed.
- R. Zeng, M. Feller, Y. Ben-David and D. Milstein, Hydrogenation and Hydrosilylation of Nitrous Oxide Homogeneously Catalyzed by a Metal Complex, J. Am. Chem. Soc., 2017, 139, 5720–5723 CrossRef CAS.
- R. Zeng, M. Feller, Y. Diskin-Posner, L. J. W. Shimon, Y. Ben-David and D. Milstein, CO Oxidation by N2O Homogeneously Catalyzed by Ruthenium Hydride Pincer Complexes Indicating a New Mechanism, J. Am. Chem. Soc., 2018, 140, 7061–7064 CrossRef CAS PubMed.
- F. Le Vaillant, A. Mateos Calbet, S. Rodriguez-Pelayo, E. J. Reijerse, S. Ni, J. Busch and J. Cornella, Catalytic Synthesis of Phenols with Nitrous Oxide, Nature, 2022, 604, 677–683 CrossRef CAS PubMed.
- S. Ni, F. Le Vaillant, A. Mateos-Calbet, R. Martin and J. Cornella, Ni-Catalyzed Oxygen Transfer from N2O onto sp3-Hybridized Carbons, J. Am. Chem. Soc., 2022, 144, 18223–18228 CrossRef CAS PubMed.
- J. Bermejo, I. Ortega-Lepe, L. L. Santos, N. Rendón, J. López-Serrano, E. Álvarez and A. Suárez, Nitrous Oxide Activation by Picoline-Derived Ni–CNP Hydrides, Chem. Commun., 2024, 60, 1575–1578 RSC.
- I. Ortega-Lepe, P. Sánchez, L. L. Santos, P. Lara, N. Rendón, J. López-Serrano, V. Salazar-Pereda, E. Álvarez, M. Paneque and A. Suárez, Catalytic Nitrous Oxide Reduction with H2 Mediated by Pincer Ir Complexes, Inorg. Chem., 2022, 61, 18590–18600 CrossRef CAS PubMed.
- J. Bermejo, L. L. Santos, E. Álvarez, J. López-Serrano and A. Suárez, Oxidation of Alcohols to Carboxylates with N2O Catalyzed by Ruthenium(II)-CNC Complexes, ACS Catal., 2025, 15, 11530–11543 CrossRef CAS.
- J. Gwak, S. Ahn, M.-H. Baik and Y. Lee, One Metal is Enough: A Nickel Complex Reduces Nitrate Anions to Nitrogen Gas, Chem. Sci., 2019, 10, 4767 RSC.
- X. Chen, H. Wang, S. Du, M. Driess and Z. Mo, Deoxygenation of Nitrous Oxide and Nitro Compounds Using Bis(N-Heterocyclic Silylene)Amido Iron Complexes as Catalysts, Angew. Chem., Int. Ed., 2022, 61, e202114598 CrossRef CAS PubMed.
- T. M. Hood, R. S. C. Charman, D. J. Liptrot and A. B. Chaplin, Copper(I) Catalysed Diboron(4) Reduction of Nitrous Oxide, Angew. Chem., Int. Ed., 2024, e202411692 CAS.
- J. S. Stanley, X. S. Wang and J. Y. Yang, Selective Electrocatalytic Reduction of Nitrous Oxide to Dinitrogen with an Iron Porphyrin Complex, ACS Catal., 2023, 13, 12617–12622 CrossRef CAS.
- J. L. Martinez, J. E. Schneider, S. W. Anferov and J. S. Anderson, Electrochemical Reduction of N2O with a Molecular Copper Catalyst, ACS Catal., 2023, 13, 12673–12680 CrossRef CAS.
- A. W. Beamer and J. A. Buss, Surface-like NOx Reduction at an Atomically-Precise Tricopper Cluster, Angew. Chem., Int. Ed., 2025, 64, e202424772 CrossRef CAS PubMed.
- N. Curado, C. Maya, J. López-Serrano and A. Rodríguez, Boryl-Assisted Hydrogenolysis of a Nickel–Methyl Bond, Chem. Commun., 2014, 50, 15718–15721 RSC.
- P. Ríos, J. Borge, F. Fernández de Córdoba, G. Sciortino, A. Lledós and A. Rodríguez, Ambiphilic Boryl Groups in a Neutral Ni(II) Complex: A New Activation Mode of H2, Chem. Sci., 2021, 12, 2540–2548 RSC.
- A. P. Deziel, M. R. Espinosa, L. Pavlovic, D. J. Charboneau, N. Hazari, K. H. Hopmann and B. Q. Mercado, Ligand and Solvent Effects on CO2 Insertion into Group 10 Metal Alkyl Bonds, Chem. Sci., 2022, 13, 2391–2404 RSC.
- We believe that the low catalytic activity observed when using THF and CH2Cl2 is related to the limited stability of the catalyst in both solvents. In CH2Cl2, nickel hydride or nickel boryl species are likely converted into nickel chloride species, which are catalytically inactive. In the case of THF, we attribute the low stability of the catalyst to the presence of traces amount of water in the solvent.
- We have carried out the reaction of nickel hydride 3 with 10 equiv. of HBpin (Fig. S29 and 30 in the SI). In the 31P{1H} NMR no significant changes were observed; however, in the 1H NMR spectra the hydride resonance of 3 (−1.7 ppm, triplet, JP–H = 34.8 Hz) disappeared upon mixing 3 with HBpin. These observations suggest that 3 and HBpin might reversibly form an adduct in which a hydride bridges both Ni and B atoms, with the equilibrium shifted toward the left side, favoring the formation of 3. Additionally, when this mixture is exposed to N2O, the formation of O(Bpin)2 is observed with generation of H2.
- The activation of N2O mediated by boryl complex 4a was also pursued, but very high energy barriers (≥32 kcal mol−1) were obtained in all cases. See SI for more details.
- P. Ríos, A. Rodríguez and J. López-Serrano, Mechanistic Studies on the Selective Reduction of CO2 to the Aldehyde Level by a Bis(phosphino)boryl (PBP)-Supported Nickel Complex, ACS Catal., 2016, 6, 5715 CrossRef.
- C. Adamo and V. Barone, Toward Reliable Density Functional Methods without Adjustable Parameters: The PBE0 Model, J. Chem. Phys., 1999, 110, 6158–6170 CrossRef CAS.
- M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision E.01, Gaussian, Inc., Wallingford, CT, 2013 Search PubMed.
- S. Grimme, J. Antony, S. Ehrlich and H. Krieg, A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H–Pu, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed.
- A. V. Marenich, C. J. Cramer and D. G. Truhlar, Universal Solvation Model Based on Solute Electron Density and a Continuum Model of the Solvent, J. Phys. Chem. B, 2009, 113, 6378–6396 CrossRef CAS.
-
(a) W. J. Hehre, R. Ditchfield and J. A. Pople, Self–Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian–Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules, J. Chem. Phys., 1972, 56, 2257–2261 CrossRef CAS;
(b) P. C. Hariharan and J. A. Pople, The Influence of Polarization Functions on Molecular Orbital Hydrogenation Energies, Theor. Chim. Acta, 1973, 28, 213–222 CrossRef CAS;
(c) M. M. Francl, W. J. Pietro, W. J. Hehre, J. S. Binkley, M. S. Gordon, D. J. DeFrees and J. A. Pople, Self–Consistent Molecular Orbital Methods. XXIII. A Polarization–Type Basis Set for Second–Row Elements, J. Chem. Phys., 1982, 77, 3654–3665 CrossRef CAS.
- M. Dolg, U. Wedig, H. Stoll and H. Preuss, Energy–Adjusted Ab Initio Pseudopotentials for the Rare Earth Elements, J. Chem. Phys., 1987, 86, 866–872 CrossRef CAS.
- A. W. Ehlers, M. Boehme, S. Dapprich, A. Gobbi, A. Höllwarth, V. Jonas, K. F. Köhler, R. Stegmann, A. Veldkamp and G. Frenking, A Set of f–Polarization Functions for Pseudo–Potential Basis Sets of the Transition Metals Sc–Cu, Y–Ag, and La–Au, Chem. Phys. Lett., 1993, 208, 111–114 CrossRef CAS.
-
(a) F. Weigend and R. Ahlrichs, Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy, Phys. Chem. Chem. Phys., 2005, 7, 3297–3305 RSC;
(b) F. Weigend, Accurate Coulomb-Fitting Basis Sets for H to Rn, Phys. Chem. Chem. Phys., 2006, 8, 1057–1065 RSC.
- S. Grimme, Supramolecular Binding Thermodynamics by Dispersion-Corrected Density Functional Theory, Chem. – Eur. J., 2012, 18, 9955–9964 CrossRef CAS PubMed.
- G. Luchini, J. V. Alegre-Requena, I. Funes-Ardoiz and R. S. Paton, GoodVibes: Automated Thermochemistry for Heterogeneous Computational Chemistry Data, F1000Research, 2020, 9, 291 Search PubMed.
- Chemcraft – Graphical Software for Visualization of Quantum Chemistry Computations. Available online: https://www.chemcraftprog.com.
|
| This journal is © the Partner Organisations 2026 |
Click here to see how this site uses Cookies. View our privacy policy here.