DOI:
10.1039/D5RA05545H
(Paper)
RSC Adv., 2025,
15, 35961-35970
Conformational constraints in [Ni(PR2NR′2)2] complexes for tuning H2 production and oxidation: a DFT-based ligand design study
Received
30th July 2025
, Accepted 18th September 2025
First published on 29th September 2025
Abstract
[Ni(PR2NR′2)2] complexes catalyze H2 production or oxidation depending on the nature of the two bidentate P2N2 ligands. Ligand design plays a crucial role in determining the reaction direction and catalytic properties. In this study, density functional theory (DFT) calculations were performed to analyze the structural and energetic properties of Ni complexes with (R, R′) = (H, H), (Me, Me), (Cy, Me), (tBu, Me), (CF3, H), and (NH2, H). Based on the structural features, relative stabilities of the NiII and Ni0 complexes, and their energy profiles, these Ni complexes were classified into three groups, I–III. In group I, the reaction is slightly exothermic toward H2 production. In group II, the presence of tBu and CF3 groups introduces steric hindrance, forcing the Ni complexes into a tetrahedral conformation. This geometric constraint destabilizes the product state in the NiII oxidation state, shifting the thermodynamics toward H2 oxidation. Conversely, destabilizing the reactant state in the Ni0 oxidation state can be achieved using a tetradentate ligand in which the two bidentate P2N2 ligands are connected by trimethylene, –(CH2)3–, units at the P atoms. This ligand, classified as group III, maintains a square planar conformation, rendering H2 production highly exothermic. These findings align with experimental observations of similar complexes and underscore the importance of ligand geometry and substituent effects.
1. Introduction
Hydrogen (H2) production and oxidation through catalytic processes have attracted significant interest as key technologies for alternative renewable energy systems intended to replace conventional fossil fuel-based systems.1–4 Ni diphosphine complexes of the form [Ni(PR2NR′2)2] (PR2NR′2 = 1,5-(di-R′)-3,7-(di-R)-1,5-diaza-3,7-diphosphacyclooctane, see Scheme 1a) have been developed.5–12 A unique feature of these complexes is the presence of two layers of coordination spheres, inspired by natural hydrogenase enzymes.7–9 The first layer is the Ni redox center, which serves as a hydride donor, while the second layer involves N atoms in pendant amines on the phosphine ligands, functioning as proton donors or acceptors. These two spheres operate cooperatively to deliver protons and enable H–H bond formation and cleavage in H2 molecules. Some Ni complexes9,13 have demonstrated exceptionally high turnover frequencies, exceeding 106 s−1, and have continued to attract attention for the development of molecular photocatalysts14,15 and electrocatalysts that promote H2 production6,13,16–18 and oxidation.19
 |
| Scheme 1 (a) Structure of Ni-diphosphine complexes [Ni(PR2NR′2)2]2+ with substituents R and R′ at the phosphorus and nitrogen atoms, respectively. Each Ni complex is denoted as (R, R′), and the combinations of substituents (R, R′) used in this study are (H, H), (Me, Me), (Cy, Me), (tBu, Me), (CF3, H), and (NH2, H). (b) A tetradentate [m3, Me] complex, in which the –(CH2)3– unit connects P1 to P3 and P2 to P4. (c) Definition of the dihedral angle (α) between planes P1 and P2, and the P–Ni–P angle (β). | |
The mechanistic details and design concepts of the Ni complexes have been investigated both experimentally10,20,21 and theoretically.21–23 The controlling factors for H2 production and oxidation were examined from a thermodynamic perspective.10 Two key parameters—hydricity (hydride donor ability)24,25 and acidity (proton donor ability)—were analyzed in relation to ligand design. The reaction mechanism was studied using spectroscopic and electrochemical measurements6,26,27 as well as theoretical calculations.21–23 In a density functional theory (DFT) study, conformational stabilities and protonation states were explored using ab initio molecular dynamics (MD) trajectories, and a plausible reaction pathway was identified.23
Another mechanistic aspect of H2 production and oxidation by Ni complexes is two-state reactivity,28,29 which involves a spin-forbidden reaction pathway enabled by intersystem crossing (ISC). Valence d-electrons in transition metals frequently facilitate ISC between potential energy surfaces of different spin multiplicities.28–31 Additionally, specific ligand coordination modes can generate electronic states that permit spin flipping during the reaction. Previous theoretical studies32,33 have shown that, in the active site of [NiFe] hydrogenase—which inspired the development of DuBois's Ni complexes,7–9 ISC can occur during the H2 binding process. The torsional angle between S–Ni–S planes was proposed as a reaction coordinate for ISC. However, the role of two-state reactivity in the H2 production and oxidation processes of DuBois's complexes has not yet been explored.
This study investigates the role of substituents in Ni diphosphine complexes [Ni(PR2NR′2)2]2+ where R and R′ are located on the phosphorus and nitrogen atoms, respectively, and which catalyze H2 production and oxidation. By introducing bulkier substituents at the R position, the dihedral angle between the P–Ni–P planes becomes tunable, influencing the relative stability of the singlet and triplet states and the possibility of ISC along the reaction pathway. Based on these findings, ligand design guidelines favorable for H2 production and oxidation are proposed.
2. Computational details
All structures (reactants, intermediates, and transition states) were fully optimized without symmetry constraints using DFT with the B3LYP34 exchange-correlation functional and Grimme's D3 (ref. 35) dispersion correction (B3LYP-D3). For Ni, the Stuttgart–Dresden relativistic effective core potential (ECP) and corresponding basis set, SDD,36 was used. For all other atoms, the 6-31G(d,p) basis set was employed for geometry optimization, normal mode analysis, and intrinsic reaction coordinate (IRC) calculations, while the 6-311++G(2d,p) basis set was used for single-point energy calculations. Hereafter, the SDD combined with 6-31G(d,p) and SDD combined with 6-311++G(2d,p) are referred to as BS1 and BS2, respectively. Basis set convergence for energy and structural parameters was evaluated using the B3LYP-D3/BS2 results as the reference. The results, presented in Section S1 of the SI, show that the energy deviations were within 0.4 kcal mol−1 (Table S1). The stability of the self-consistent field (SCF) solutions was confirmed through stability analysis.37 Zero-point vibrational energy corrections at the BS1 level were added to the potential energy. The choice of DFT exchange–correlation functionals was validated by comparison with single-point CCSD(T) calculations for the (Me, Me) complex in a previous study.38 See Table S2 in Section S2 of SI for details. Transition states were confirmed by identifying a single imaginary frequency in harmonic frequency analysis. IRC calculations were performed to verify the reliability of the reaction pathways. Structure optimizations of energy minima, transition states, minimum energy crossing points (MECPs), normal mode analyses, and IRC calculations were conducted using the GRRM23 program,39,40 which interfaces with Gaussian16 (ref. 41) for computing potential energy, energy gradients, and Hessians. Solvation effects were included using the SMD method42 with parameters for acetonitrile. Spin–orbit coupling (SOC) matrix elements between spin states were calculated using the MolSOC program.43,44
3. Results and discussion
3.1 Relative stability of singlet and triplet states of the NiII and Ni0 complex with different substituents
The Ni complex, [NiII(PR2NR′2)2]2+, with substituent groups at the R and R′ positions is referred to as (R, R′) as shown in Scheme 1a. This NiII complex corresponds to the C and P intermediates in the reaction mechanism. Six complexes—(H, H), (Me, Me), (Cy, Me), (tBu, Me), (CF3, Me), and (NH2, Me)—were investigated. Based on the results, an additional complex was examined featuring a tetradentate ligand in which two bidentate ligands are connected at the phosphorus atoms by a –(CH2)3– unit. This complex is denoted as the [m3, Me] complex. Following the findings of a previous study by Raugei et al.,23 possible conformations of the (H, H) complex were optimized (Fig. S1 and Table S3). Among the optimized structures, the lowest-energy conformation—schematically shown in Scheme 1a—was selected. In this conformation, the two six-membered (Ni–P–C–N–C–P) rings in the bidentate ligands adopt boat and chair conformations, and energy of the conformation is 2.5 kcal mol−1 more stable than the other conformers, in qualitative agreement with the earlier study.23 For the other complexes, the relative stabilities of the conformers were also evaluated and are listed in Table S4. The (Me, Me), (NH2, Me), (CF3, H) and [m3, Me] complexes adopt boat–chair conformations, while in the (Cy, Me) and (tBu, Me) complexes, the boat–boat conformation is slightly more stable.
The optimized structures of the [NiII(PR2NR′2)2]2+ complexes in both singlet and triplet states are presented in Fig. S2. In the reaction pathway, this is a product state (P) after reductive elimination of H2 is completed. The relative energies ΔE (kcal mol−1), dihedral angles (°), and P–Ni–P angles (°) for the singlet (S) and triplet (T) ground states are summarized in Table 1. The definitions of the dihedral and P–Ni–P angles are provided in Scheme 1b. Structure optimization revealed that the NII (H, H), (Me, Me), (Cy, Me) adopt square planar geometries in the singlet state and tetrahedral geometries in the triplet state. As a representative example, the structures of the (Me, Me) complex in singlet states is shown in Fig. 1a. In contrast, the (tBu, Me) complex exhibits an exceptional behavior: as shown in Fig. 1b, its optimized structure is tetrahedral in both singlet and triplet states. The dihedral angle (α), reported in Table 1, supports this observation. Only the (tBu, Me) complex, with a bulky tBu group at each P atom, exhibits a dihedral angle of 72.8°, indicating a tetrahedral configuration. Complexes with H, Me, or Cy substituents exhibit dihedral angles ranging from 0° to 30.8°, indicative of nearly square planar geometries. In the triplet state, the dihedral angles range from 74.2° to 84.3°, in clear contrast to those in the singlet state. The unusually large dihedral angle in the singlet state of the (tBu, Me) complex arises from steric repulsion between the tBu groups.
Table 1 Relative energies ΔE (kcal mol−1), dihedral angles α (°), and P–Ni–P angles β (°) of the [NiII(PR2NR′2)2]2+ complex in singlet (S) and triplet (T) states (B3LYP-D3//BS1 results)
(R, R′) |
ΔE (kcal mol−1) |
α (°) |
β (°) |
S |
T |
S |
T |
Sa |
T (P1–Ni–P3)/(P2–Ni–P4) |
P1–Ni–P3 and P2–Ni–P4 angles are equal to the first decimal place in the singlet state. |
(H, H) |
0.0 |
11.4 |
0.0 |
81.6 |
96.2 |
164.9/106.3 |
(Me, H) |
0.0 |
5.4 |
24.4 |
84.3 |
99.4 |
107.9/107.8 |
(Me, Me) |
0.0 |
5.2 |
26.5 |
78.6 |
100.5 |
104.9/104.9 |
(Cy, Me) |
0.0 |
7.4 |
30.8 |
74.2 |
104.0 |
117.3/117.3 |
(tBu, Me) |
14.9 |
0.0 |
72.8 |
78.5 |
113.8 |
142.6/111.4 |
 |
| Fig. 1 Structures of the (a) NiII (Me, Me) and (b) (tBu, Me) complexes in singlet states (atomic distances are in Å). | |
The structural constraint in the NiII (tBu, Me) complex significantly affects the relative stability of the singlet and triplet states. In the (H, H), (Me, Me), and (Cy, Me) complexes, the singlet state is more stable than the triplet state. In contrast, this trend is reversed in the (tBu, Me) complex, where the triplet state is 14.9 kcal mol−1 more stable than the singlet state, as shown in Table 1. Compared with the (Me, Me) complex, the relative energy difference is reversed by 20.1 kcal mol−1 (5.2 + 14.9). From a ligand design perspective, it is important to note that introducing a tBu group can reverse the singlet–triplet stability by approximately 20 kcal mol−1.
In contrast to the (tBu, Me) complex, the (Cy, Me) complex exhibits a less distorted dihedral angle (30.8°) and does not adopt a tetrahedral conformation in the singlet state. This indicates that the Cy groups can avoid steric repulsion even at relatively small dihedral angles. Each Cy group interacts with the other through a C–H moiety adjacent to the P atom. In the (tBu, Me) complex, this corresponding carbon is a tertiary carbon bearing a methyl group. Additionally, the Cy groups exhibit attractive interactions with the N atoms of the opposing ligands, with an H–N distance of 2.46 Å, as shown in Fig. S2. Therefore, the dihedral angle in the (Cy, Me) complex is governed by a balance between repulsive H–H and attractive H–N interactions, resulting in a nearly planar geometry in the singlet state.
We also calculated the structure of the Ni0 (H, H) complex, Ni0(PH2NH2)2, in singlet state for comparison. In the reaction pathway, this Ni0 complex corresponds to the R intermediate which is a resting state before accepting protons. The Ni0 center has a d10 electron configuration, and isotropic ligand coordination leads to a tetrahedral geometry. The dihedral angle (α) between the two Ni–P–P planes is 89.5°, in clear contrast to the NiII case (0.4 degrees). Constrained structure optimization with fixed α was performed, and the results are shown in Fig. 2 and S3. By twisting the dihedral angle by 45°, the NiII complex is destabilizes by approximately 9 kcal mol−1. Similarly, the 45° twist from the tetrahedral structure destabilizes the Ni0 complex by about 15 kcal mol−1. The variation in stability with structural distortion should be recognized as a valuable feature in molecular design.
 |
| Fig. 2 Potential energy curves of Ni0 and NiII (H, H) complexes. The energy reference for each complex is its fully optimized structure (α = 0.4° for NiII and α = 89.4° for Ni0). For all other structures, only the dihedral angle was fixed, while all other structural parameters were optimized. | |
3.2 Energy profiles of H2 production and oxidation
The reaction cycle is described below in the direction of H2 production. The reaction pathway for H2 production/oxidation by [Ni(PR2NR′2)2]2+ complexes has been proposed in previous studies.23,45–47 Two major mechanisms—EECC and ECEC—have been discussed, where E and C represent electron and proton transfers, respectively. Depending on the applied potential, either the EECC or ECEC mechanism may dominate. According to a kinetic study,47 the EECC pathway proceeds faster than the ECEC pathway. In this study, the EECC mechanism shown in Scheme 2 was adopted. H2 production begins with the adsorption of two protons (H1 and H2) at the pendant amine ligands (N1 and N2) of the Ni0 resting state (R), forming the diprotonated intermediate (A) with a Ni0 metal center. A Brønsted acid (protonated base, HBase+) serves as the proton source. The A intermediate then undergoes an intramolecular proton-transfer transition state (TSB1), resulting in the hydride–proton intermediate (B1), where one proton (H1) migrates from the amine (N1) to the Ni center. Following a transition state (TSB2) with a very low energy barrier, the hydride at the Ni center adopts a new conformation (B2 intermediate) favorable for H–H bond formation. This conformational change was observed only in the (H, H), (Cy, Me), and (tBu, Me) complexes. The B2 intermediate then undergoes a second intramolecular proton transfer (H2) via a transition state (TSC), producing the H2 complex (C) with a NiII center. Finally, in the product state P, the NiII complex and the generated H2 molecule are treated as isolated species. The P state accepts two electrons to regenerate the R state, followed by two proton transfers to form the A intermediate again.
 |
| Scheme 2 Reaction pathway considered in this study. | |
According to the [NiFe] hydrogenase model,32,33 a low-lying triplet state appears along the reaction pathway. Energy diagrams for both singlet and triplet states were obtained, and the results for the (H, H), (Me, Me), (Cy, Me) and (tBu, Me) complexes are summarized in Fig. 3 and S4 of the SI. For clarity, some of the triplet-state diagrams are omitted in Fig. 3. The potential energies of the singlet and triplet states are closely aligned around the TSC region. For the (H, H) complex, CCSD/6-311+G(2d,p) results indicate that the 1TSC state is 4.3 kcal mol−1 more stable than the corresponding triplet state. DFT results for the (H, H), (Me, Me), and (Cy, Me) complexes also show no intersystem crossing, as illustrated in Fig. S4. An exception case is the (tBu, Me) complex, which will be discussed later.
 |
| Fig. 3 (a) Potential energy diagram with zero-point vibrational energy correction, calculated at the B3LYP-D3/BS2//B3LYP-D3/BS1 level. The legends are consistent with those in Fig. 3a. The “×” symbol indicates the minimum-energy intersystem crossing point. The energy of the crossing point does not include zero-point energy correction. (b) Dihedral angle (α) of the Ni complex along the reaction pathway (for definition, see Fig. 2). | |
3.2.1 (H, H), (Me, Me), (Cy, Me) and (tBu, Me) cases. An overview of the energy profile and structural changes is presented in this section, while specific features of each complex are discussed in the following sections. These Ni complexes are classified into two groups. The first group includes the (H, H), (Me, Me), and (Cy, Me) complexes, in which the reaction proceeds through the singlet state. In this group, the 1A and 1(P+H2) states are nearly isoenergetic, and the H2 production process is calculated to be slightly exothermic, with reaction energies ranging from −3.2 to −5.9 kcal mol−1. The rate determining step (RDS) is either 1TSB1 (Ni–H formation) or 1TSC (reductive elimination), depending on the substituents. For the (H, H) and (Me, Me) complexes, the RDS is 1TSC, with apparent activation energies of 12.8 kcal mol−1 and 9.2 kcal mol−1, respectively. In the (Cy, Me) complexes, the RDS is 1TSB1, with an activation energy of 8.0 kcal mol−1.As a representative case, the (H, H) complex was selected, and the optimized structures of the intermediates and transition states are shown in Fig. 4. For the other complexes, qualitatively similar structures were obtained, as shown in Fig. S5(a)–(i) in SI. Structural changes are described along the H2 production pathway. The 1A state is formed after the 1R state accepts two protons at the nitrogen atoms of the pendant amines. Both the H1–Ni and H2–Ni distances are 2.76 Å. One of the protons (H1) is transferred to the Ni center via 1TSB1, during which the H1–Ni distance shortens to 1.63 Å. The structure of the 1B1 state adopts a pseudo-trigonal bipyramidal geometry. The complex then undergoes structural isomerization to reach the 1B2 state, where the H1–H2 distance becomes 1.91 Å, indicating that interaction for H–H bond formation has already begun (see Fig. 4d). The H2 molecule is formed via the 1TSC state, where the dihedral angle α between the two Ni–P–P planes decreases to below 40°. H2 formation completes at the 1C and 1(P+H2) states, where the dihedral angle approaches 90°, reflecting the characteristic structure of the NiII complex.
 |
| Fig. 4 Optimized structures of the (H, H) complex in (a) 1A, (b) 1TSB1, (c) 1B1, (d) 1B2, (e) 1TSC and (f) 1P (bond lengths are in Å). | |
The second group comprises the (tBu, Me) complex. The primary distinction from the (H, H), (Me, Me), and (Cy, Me) complexes lies in the behavior of the singlet surface. As shown in Fig. 3a, the singlet states of the (tBu, Me) complex in 1TSC, 1C, and 1(P+H2) are less stable than their corresponding triplet states by 15.8, 28.1, and 25.5 kcal mol−1, respectively. In contrast, the energy levels of the 1A, 1TSB1, 1B1, 1TSB2 and 1B2 states are comparable to those of the other complexes. This instability arises from structural distortion caused by the steric hindrance of the bulky tBu ligands. As shown in Fig. 1b, the tBu groups on the two P2N2 ligands avoid each other while coordinating to the Ni center, resulting in a tetrahedral conformation. In the 1TSC, 1C, and 1(P+H2) states, after two H+ are reduced, the metal center is oxidized to NiII. As shown in Table 1 and Fig. 2, the NiII complexes of (H, H), (Me, Me) and (Cy, Me)—which lack steric hindrance—favor a square planar structure in the singlet state. The dihedral angle α in the 1(P+H2) state is less than 40°, as seen in Fig. 3b. In contrast, the dihedral angle in the (tBu, Me) complex is 57.5°, and is constrained toward a tetrahedral conformation due to the presence of the tBu group.
As the relative stability of the singlet and triplet states reverses around the 1TSC state, intersystem crossing is expected. The intersystem crossing point was located at 29.1 kcal mol−1 (relative to the 1A state, without zero-point energy correction) between the 1B2 and 1TSC states, and calculated SOC matrix element is 44.9 cm−1 (for more details, see Table S5).
A key finding from the (tBu, Me) case is that the thermodynamics of H2 production and oxidation in Ni complexes can be tuned by imposing square planar or tetrahedral constraints through steric hindrance of the substituents. As a result, the (tBu, Me) complex is more favorable for H2 oxidation than the other complexes. Along the reaction pathway involving intersystem crossing, 3(P+H2) →→ 3TSC → 1B2 →→ 1A, H2 oxidation is exothermic, with a calculated reaction energy of −4.9 kcal mol−1. The rate-determining step is the intersystem crossing between the 3TSC and 1B2 states, with a barrier height of 24.2 kcal mol−1.
3.2.2 Controlling the dihedral angle of the Ni complex for improving H2 production capability. In the (H, H), (Me, Me), and (Cy, Me) complexes, the reaction energy for H2 production ranges from −3.2 to −5.9 kcal−1. With the introduction of tBu ligands the reaction energy increases to +4.9 kcal mol−1, making the (tBu, Me) complex thermodynamically unfavorable for H2 production but favorable for H2 oxidation. This section explores ligand design strategies to enhance the suitability of Ni complexes for H2 production.To modify the energy profile to favor exothermic H2 production, the energy level of the 1A state can be raised relative to that of the 1C state. As noted, this relative energy can be adjusted by the dihedral angle α. As shown in Fig. 2, constraining the dihedral angle to 45 degrees effectively raises the energy level of the Ni0 complex by approximately +15 kcal mol−1. The 1A state is in Ni0 oxidation state, which prefers a tetrahedral geometry. Based on this, the Ni0 complex was forced into a square planar geometry by linking the two bidentate phosphine ligands. In the [Ni(PR2NR′2)2] structure, a trimethylene, –(CH2)3–, unit was introduced at the R position to connect the two P2N2 ligands. A methyl (Me) group was used for the R′ substituent. This new complex is referred to as [m3, Me] and is shown in Scheme 1b.
The potential energy profiles for H2 production/oxidation by the [m3, Me] complex in the singlet and triplet states were calculated and are shown in Fig. 3a (singlet state) and Fig. S4 (both singlet and triplet states). The optimized structures of the intermediates and transition states in singlet states are presented in Fig. 5 and S5f. H2 production proceeds along the reaction pathway in the singlet state, with an energy barrier of 4.5 kcal mol−1 at 1TSB1. This is the lowest barrier among all Ni complexes examined in this study. Following the 1B1 intermediate, the complex passes through 1TSC with a negligible barrier and proceeds to 1C and 1P to form the H–H bond. The reason for the low-barrier heights of the transition states is in the structural similarities between 1A and 1TSB1 and between 1B1 and 1TSC. Especially, the H–H distance in 1B1 is 1.54 Å and is close to that in 1TSC (1.09 Å). This is in clear contrast to those in the (H, H) complex, where the H–H distance in 1B2 is 1.91 Å as shown in Fig. 4. The tetradentate P4N4 ligand not only constrains square planar geometry but also imposes the structures of 1A and 1B1 close to that of 1TSC. The reaction energy for H2 production is −40.6 kcal mol−1, making it highly exothermic and thermodynamically the most favorable among the studied Ni complexes. The structural constraint effectively raises the energy levels of the 1A and 1B1 states. The dihedral angle of the [m3, Me] complex along the reaction pathway is also shown in Fig. 3b. The trend in dihedral angle differs markedly from that of the other complexes, ranging from 31.7° to 0.3° across the 1A, 1TSB1, 1B1, 1TSC, 1C, and 1P states.
 |
| Fig. 5 Optimized structures of the [m3, Me] complex in the singlet state (bond lengths are in Å). | |
Previous studies reported experimental results for Co48,49 and Ni13 complexes bearing tetradentate phosphine ligands. Wiedner and co-workers synthesized Co complexes with P4N2-type tetradentate phosphine ligands48 and successfully achieved H2 production with a high turnover frequency (1.8 × 104 s−1) among the highest for molecular catalysts.49 This turnover frequency was further enhanced in a tetradentate Ni complex, reaching up to 1.6 × 106 s−1.13 Thus, the proposed molecular design concept is supported both theoretically by proof-of-concept DFT calculations and experimentally by the high catalytic activity of tetradentate ligands.
Notably, previous studies demonstrated a linear correlation between the dihedral angle (α)50 and the hydride donor ability
of [HM(LH)]2+, where M and L represent the metal and ligand, respectively.51
A smaller dihedral angle increases donor ability, contributing to the exothermicity of the elementary step from the 1B to 1C states. However,
is only one factor in explaining the reactivity. The energy difference between the 1B and 1(P+H2) states also includes the energy required for proton release ([HM(LH)]2+ → [HML]+ + H+) and H2 formation (H+ + H− → H2).49,52 Moreover,
is not related to the energy difference between the 1A and 1B1 states. The results for the [m3, Me] complex indicate that enforcing a square planar geometry in the 1A and 1B1 states is a key strategy for promoting efficient H2 production.
3.2.3 Classification of the Ni complexes. The five complexes discussed above are classified into three categories (I–III) based on ligand rigidity and the reaction energy associated with H2 production/oxidation. In Fig. 6, the reaction energy for H2 production is plotted against the change in dihedral angle from the A to P+H2 states. The dashed horizontal lines in Fig. 6, connecting the A to P+H2 states for each complex, represent the change in dihedral angle along the reaction pathway. Group I includes the (H, H), (Me, Me), and (Cy, Me) complexes, which exhibit slightly exothermic reaction energies for H2 production. These ligands are relatively less bulky and flexible enough to accommodate structural changes between tetrahedral and square planar geometries depending on the oxidation state of the Ni center. This is reflected by the characteristic long horizontal lines observed in Fig. 6. Group II consists of the (tBu, Me) complex, which thermodynamically favors H2 oxidation. The dihedral angles remain around 90° due to steric hindrance, which restricts structural rearrangement of the Ni complex. Group III comprises the [m3, Me] complex featuring a tetradentate ligand. The dihedral angles range from 0° to 30°, indicating strong structural enforcement toward planar configurations. This complex shows a highly exothermic profile for H2 production.
 |
| Fig. 6 Reaction energies for the H2 production (from A to P+H2 state) is plotted against the dihedral angles of the A and P+H2 states. Filled and open symbols represent the P+H2 and A states, respectively. The dashed line connects the A to P+H2 states, indicating the change in dihedral angle during the reaction. Single-point B3LYP-D3/BS2 calculations with zero-point vibrational correction were used to determine the reaction energies. For the (tBu, Me) and (CF3, H) complexes, pathways via the intersystem crossing point were used, and the reaction energy is defined as the energy difference between the triplet 3(P+H2) state and the singlet 1A state. Numerical data are provided in Table S6 of the SI. | |
The ligands were selected to control the dihedral angles of the Ni complexes through structural constraints. As shown in Fig. 2, these dihedral angles clearly influence the relative stability of the tetrahedral and square planar geometries in the Ni0 and NiII complexes. However, additional factors should be considered in ligand design. In particular, the electron-withdrawing and electron-donating properties of the substituents may influence the trends observed in Fig. 6. To examine these effects, the (CF3, H) and (NH3, H) complexes were analyzed. The (CF3, H) complex was compared with the (tBu, Me) complex, as the CF3 group is similarly bulky but more electron-withdrawing than the tBu group. The (NH2, H) complex was compared with the group I complexes to assess the impact of the electron-donating nature of the NH2 group.
The results are also presented in Fig. 6. The CF3 group clearly constrains the structure to a tetrahedral geometry. The dihedral angles in the 1A and 3(P+H2) states are 88.6° and 79.2°, respectively. The reaction energy for H2 oxidation is exothermic by −26.8 kcal mol−1, which is significantly greater than that of the (tBu, Me) complex (−4.9 kcal mol−1). This difference arises because the triplet state energy levels in the 3C and 3(P+H2) states are higher for the CF3-substituted complex than for the (tBu, Me) complex, as shown in Fig. 3a. Energy diagram of the (CF3, H) complex shown in Fig. 3b also indicates that the H2 oxidation is feasible. As in the (tBu, Me) case, the initial state of the (CF3, H) complex is in triplet state, 3(P+H2) state. In TSC, singlet and triplet states are nearly degenerate, and the energy difference is only 1.0 kcal mol−1. Minimum energy ISC point was located at 39.4 kcal mol−1, which is higher than both 1TSC and 3TSC by 3.3 and 1.7 kcal mol−1, respectively. This ISC point is at 12.6 kcal mol−1 from the 3(P+H2) state and by 13.9 kcal mol−1 smaller than that of the (tBu, Me) case. Structure of the ISC point is halfway between 1B2 and 3TSC as shown in Fig. S5g. Calculated SOC is 26.1 cm−1, which is in the same order as that of the (tBu, Me) case (44.9 cm−1). After ISC, the reaction should go through the singlet surface until two protons are formed at 1A state. The electron-withdrawing nature of the CF3 group enhances the H2 oxidation capability of the Ni complex. In contrast, the (NH2, H) complex exhibits behavior similar to that of the group I complexes in both structural changes and reaction energy toward H2 production.
As the reaction barrier in the reaction pathway would block the H2 reduction and oxidation reaction, we summarized the reaction barrier of the Ni complexes in Table S7. The result show that the activation barrier would be small enough not to block the H2 reduction in the groups I and III complexes and the H2 oxidation in the group II complexes. One exception is the H2 oxidation by the (tBu, Me) case in which calculated activation barrier is 24.2 kcal mol−1. Those for the other complexes are less than 13 kcal mol−1, illustrating the catalytic ability of the Ni complexes.
The final part of this section discusses the influence of acids and bases on the reaction energy. In H2 oxidation experiments, bases such as triethylamine, NEt3, are often used as proton acceptor.5,19,24 H2 production experiments commonly employ [DMF(H)][OTf] as a proton donor.6,13,16 In this study, NEt3 and [DMF(H)][OTf] were adopted as representative base and acid, respectively. As shown in Table 2, the acid and base strongly promote H2 production and oxidation, respectively. In group I, the acid and base play a decisive role. The energy profile of the (Me, Me) complex becomes exothermic toward H2 production and oxidation in the presence of [DMF(H)][OTf] and NEt3, respectively. In contrast, the (CF3, H) complex in group II is exothermic toward H2 oxidation under both NEt3 and [DMF(H)][OTf]. The [m3, Me] complex in group III favors H2 production regardless of the presence of acid or base. These results indicate that the properties of the complexes can be tuned through ligand design to surpass the influence of external acids and bases.
Table 2 Reaction energies including the proton donor in H2 production and the proton acceptor in H2 oxidation (values are given in kcal mol−1)
|
ΔE (H2 oxidation)a |
ΔE (H2 production)a |
H+ acceptor: NEt3 |
H+ donor: [DMF(H)][OTf] |
B3LYP-D3 |
ωB97XD |
B3LYP-D3 |
ωB97XD |
Reaction energy with zero-point vibration energy correction. For H2 oxidation, ΔE = E(R) + 2E(H+NEt3) − E(P) − E(H2) − 2E(NEt3). For H2 production, ΔE = E(P) + E(H2) + 2E([DMF][OTf]−) − E(R) − 2E([DMF(H)][TOf]). |
(Me, Me) |
−13.7 |
1.5 |
−14.9 |
−25.2 |
(tBu, Me) |
−20.4 |
−5.2 |
−8.2 |
−18.5 |
(CF3, H) |
−61.6 |
−44.1 |
33.0 |
20.4 |
[m3, Me] |
26.8 |
43.3 |
−55.4 |
−67.0 |
4. Conclusion
The Ni complexes developed by DuBois et al.5–12 have demonstrated strong potential for electrocatalytic H2 production and oxidation. As these two reactions proceed in opposite directions, ligand design plays a critical role in controlling their directionality. In this study, DFT calculations were performed for [Ni(PR2NR′2)2] complexes with (R, R′) = (H, H), (Me, Me), (Cy, Me), (tBu, Me), (CF3, H) and (NH2, H) to investigate how ligands influence H2 production and oxidation activity. Based on the analysis of conformational changes during the reaction, a [m3, Me] complex featuring a P4N4-type tetradentate ligand is proposed. A proof-of-concept calculation was conducted to evaluate its energy profile.
To understand fundamental properties, the structures of the NiII and Ni0 complexes were examined. In the singlet state, the NiII complex adopts a square planar conformation, whereas the Ni0 complex prefers a tetrahedral conformation. A notable exception is the NiII (tBu, Me) complex, which exhibits a tetrahedral structure in the singlet state due to steric repulsion between the bulky tBu groups, forcing the complex into a tetrahedral geometry. This steric hindrance in the NiII state selectively destabilizes the energy levels of the TSC, C, and P states in the singlet state. As a result, the triplet state becomes the ground state for these intermediates in the NiII (tBu, Me) complex. This finding provides a rationale for designing a tetradentate ligand.
The energy diagrams of the (H, H), (Me, Me), and (Cy, Me) complexes are slightly exothermic—by −3.2 to −5.9 kcal mol−1—toward H2 production, whereas that of the (tBu, Me) complex is exothermic by −4.9 kcal mol−1 toward H2 oxidation. Structural analysis suggests a ligand design in which two bidentate PR2NR′2 ligands are connected by trimethylene –(CH2)3– units to enforce a square planar geometry in the Ni0 complex. This ligand design effectively destabilizes the 1A, 1TSB1 and 1B1 states of the [m3, Me] complex, shifting the reaction energy for H2 production to a highly exothermic value of −40.6 kcal mol−1. The calculated barrier height at the rate-determining 1TSB1 state is 4.5 kcal mol−1. These findings are consistent with experimental studies on related systems13 and offer valuable guidance for the design of high-performance catalysts.
To verify this understanding, an electron-withdrawing CF3 group and an electron-donating NH2 group were introduced. The reaction energy of the (NH2, H) complex remains consistent with the trend observed in group I complexes. The sterically bulky CF3 group enforces a tetrahedral conformation and destabilizes the 3P state, increasing the reaction energy for H2 oxidation to −26.8 kcal mol−1.
These Ni complexes were classified into three groups. Group I included the (H, H), (Me, Me), (Cy, Me), and (NH2, H) complexes. Group II comprised the (tBu, Me) and (CF3, H) complexes, which exhibited exothermic behavior toward H2 oxidation. The [m3, Me] complex belonged to group III and favored H2 production. The effect of proton donors and acceptors on H2 production and oxidation was examined by introducing [DMF(H)][OTf] and NEt3, respectively. The presence of acid and base strongly influences the direction of the reaction toward H2 production or oxidation, respectively. However, our calculations indicate that group II Ni complexes—particularly the (CF3, H) complex—can thermodynamically favor H2 oxidation even in the presence of acid. In contrast, the [m3, Me] complex prefers H2 production even under the presence of base.
The results of this study show that the conformation of the Ni complex can effectively and selectively shift specific energy levels along the reaction pathway. Designing the ligand structure is essential for controlling the directionality of H2 production or oxidation.
Conflicts of interest
There are no conflicts to declare.
Data availability
The data supporting this article (basis-sets and DFT-functional dependences, structures of Ni complexes, potential energy of the Ni complex in singlet and triplet states, spin-orbit coupling constants at intersystem crossing points, summary of reaction barrier and reaction energy, and atomic coordinates) are provided in the supplementary information (SI). Supplementary information is available. See DOI: https://doi.org/10.1039/d5ra05545h.
Acknowledgements
This study was supported by JSPS KAKENHI (grant numbers JP23K17898, JP24K08346, JP24K17554, and JP25H01266). Additional support was provided by the Photoexcitonix Project, the Junior Scientist Promotion Program at Hokkaido University, the Joint Usage and Research Center for Catalysis (JURCC), and the Integrated Research Consortium on Chemical Sciences (IRCCS). Part of the computations was performed at the Research Center for Computational Science, Okazaki, Japan (project: 25-IMS-C002) and Center for Computational Materials Science, Institute for Materials Research, Tohoku University for the use of MASAMUNE-II (project No. 202412-SCKXX-0007). Financial support was also provided by the National Science, Research, and Innovation Fund (NSRF). SJ gratefully acknowledges support from Program Management Unit for Human Resources & Institutional Development, Research and Innovation (grant number B42G680032). M. R. acknowledges the MEXT Promotion of the Development of a Joint Usage/Research System Project: Coalition of Universities for Research Excellence (CURE) Program [Grant Number JPMXP1323015474].
References
- W. Lubitz and W. Tumas, Hydrogen: An Overview, Chem. Rev., 2007, 107, 3900–3903 CrossRef CAS PubMed.
- S. Suman, Hybrid nuclear-renewable energy systems: A review, J. Cleaner Prod., 2018, 181, 166–177 CrossRef.
- T. da Silva Veras, T. S. Mozer, D. da Costa Rubim Messeder dos Santos and A. da Silva César, Hydrogen: Trends, production and characterization of the main process worldwide, Int. J. Hydrogen Energy, 2017, 42, 2018–2033 CrossRef CAS.
- D. Monga, N. P. Shetti, S. Basu and R. R. Kakarla, Recent advances in various processes for clean and sustainable hydrogen production, Nano-Struct. Nano-Objects, 2023, 33, 100948 CrossRef CAS.
- C. J. Curtis, A. Miedaner, R. Ciancanelli, W. W. Ellis, B. C. Noll, M. Rakowski DuBois and D. L. DuBois, [Ni(Et2PCH2NMeCH2PEt2)2]2+ as a Functional Model for Hydrogenases, Inorg. Chem., 2003, 42, 216–227 CrossRef CAS PubMed.
- A. D. Wilson, R. H. Newell, M. J. McNevin, J. T. Muckerman, M. Rakowski DuBois and D. L. DuBois, Hydrogen Oxidation and Production Using Nickel-Based Molecular Catalysts with Positioned Proton Relays, J. Am. Chem. Soc., 2006, 128, 358–366 CrossRef CAS PubMed.
- M. Rakowski DuBois and D. L. DuBois, The roles of the first and second coordination spheres in the design of molecular catalysts for H2 production and oxidation, Chem. Soc. Rev., 2009, 38, 62–72 RSC.
- M. Rakowski Dubois and D. L. Dubois, Development of Molecular Electrocatalysts for CO2 Reduction and H2 Production/Oxidation, Acc. Chem. Res., 2009, 42, 1974–1982 CrossRef CAS PubMed.
- M. L. Helm, M. P. Stewart, R. M. Bullock, M. R. DuBois and D. L. DuBois, A Synthetic Nickel Electrocatalyst with a Turnover Frequency Above 100,000 s−1 for H2 Production, Science, 2011, 333, 863–866 CrossRef CAS PubMed.
- D. L. DuBois and R. M. Bullock, Molecular Electrocatalysts for the Oxidation of Hydrogen and the Production of Hydrogen – The Role of Pendant Amines as Proton Relays, Eur. J. Inorg. Chem., 2011, 2011, 1017–1027 CrossRef.
- W. J. Shaw, M. L. Helm and D. L. DuBois, A modular, energy-based approach to the development of nickel containing molecular electrocatalysts for hydrogen production and oxidation, Biochim. Biophys. Acta, Bioenerg., 2013, 1827, 1123–1139 CrossRef CAS.
- D. L. DuBois, Development of Molecular Electrocatalysts for Energy Storage, Inorg. Chem., 2014, 53, 3935–3960 CrossRef CAS PubMed.
- C. M. Klug, W. G. Dougherty, W. S. Kassel and E. S. Wiedner, Electrocatalytic Hydrogen Production by a Nickel Complex Containing a Tetradentate Phosphine Ligand, Organometallics, 2019, 38, 1269–1279 CrossRef CAS.
- M. A. Gross, A. Reynal, J. R. Durrant and E. Reisner, Versatile Photocatalytic Systems for H2 Generation in Water Based on an Efficient DuBois-Type Nickel Catalyst, J. Am. Chem. Soc., 2014, 136, 356–366 CrossRef CAS PubMed.
- Y. Zhou, S. Yang and J. Huang, Light-driven hydrogen production from aqueous solutions based on a new Dubois-type nickel catalyst, Phys. Chem. Chem. Phys., 2017, 19, 7471–7475 RSC.
- U. J. Kilgore, M. P. Stewart, M. L. Helm, W. G. Dougherty, W. S. Kassel, M. R. DuBois, D. L. DuBois and R. M. Bullock, Studies of a Series of [Ni(PR2NPh2)2(CH3CN)]2+ Complexes as Electrocatalysts for H2 Production: Substituent Variation at the Phosphorus Atom of the P2N2 Ligand, Inorg. Chem., 2011, 50, 10908–10918 CrossRef CAS PubMed.
- U. J. Kilgore, J. A. S. Roberts, D. H. Pool, A. M. Appel, M. P. Stewart, M. R. DuBois, W. G. Dougherty, W. S. Kassel, R. M. Bullock and D. L. DuBois, [Ni(PPh2NC6H4X2)2]2+ Complexes as Electrocatalysts for H2 Production: Effect of Substituents, Acids, and Water on Catalytic Rates, J. Am. Chem. Soc., 2011, 133, 5861–5872 CrossRef.
- M. P. Stewart, M.-H. Ho, S. Wiese, M. L. Lindstrom, C. E. Thogerson, S. Raugei, R. M. Bullock and M. L. Helm, High Catalytic Rates for Hydrogen Production Using Nickel Electrocatalysts with Seven-Membered Cyclic Diphosphine Ligands Containing One Pendant Amine, J. Am. Chem. Soc., 2013, 135, 6033–6046 CrossRef PubMed.
- J. Y. Yang, S. Chen, W. G. Dougherty, W. S. Kassel, R. M. Bullock, D. L. DuBois, S. Raugei, R. Rousseau, M. Dupuis and M. R. DuBois, Hydrogen oxidation catalysis by a nickel diphosphine complex with pendant tert-butyl amines, Chem. Commun., 2010, 46, 8618–8620 RSC.
- M. O’Hagan, M.-H. Ho, J. Y. Yang, A. M. Appel, M. R. DuBois, S. Raugei, W. J. Shaw, D. L. DuBois and R. M. Bullock, Proton Delivery and Removal in [Ni(PR2NR’2)2]2+ Hydrogen Production and Oxidation Catalysts, J. Am. Chem. Soc., 2012, 134, 19409–19424 CrossRef PubMed.
- M. O'Hagan, W. J. Shaw, S. Raugei, S. Chen, J. Y. Yang, U. J. Kilgore, D. L. DuBois and R. M. Bullock, Moving Protons with Pendant Amines: Proton Mobility in a Nickel Catalyst for Oxidation of Hydrogen, J. Am. Chem. Soc., 2011, 133, 14301–14312 CrossRef PubMed.
- M. Dupuis, S. Chen, S. Raugei, D. L. DuBois and R. M. Bullock, Comment on “New Insights in the Electrocatalytic Proton Reduction and Hydrogen Oxidation by Bioinspired Catalysts: A DFT Investigation”, J. Phys. Chem. A, 2011, 115, 4861–4865 CrossRef PubMed.
- S. Raugei, S. Chen, M.-H. Ho, B. Ginovska-Pangovska, R. J. Rousseau, M. Dupuis, D. L. DuBois and R. M. Bullock, The Role of Pendant Amines in the Breaking and Forming of Molecular Hydrogen Catalyzed by Nickel Complexes, Chem.–Eur. J., 2012, 18, 6493–6506 CrossRef PubMed.
- C. J. Curtis, A. Miedaner, W. W. Ellis and D. L. DuBois, Measurement of the Hydride Donor Abilities of [HM(diphosphine)2]+ Complexes (M = Ni, Pt) by Heterolytic Activation of Hydrogen, J. Am. Chem. Soc., 2002, 124, 1918–1925 CrossRef PubMed.
- D. E. Berning, B. C. Noll and D. L. DuBois, Relative Hydride, Proton, and Hydrogen Atom Transfer Abilities of [HM(diphosphine)2]PF6 Complexes (M = Pt, Ni), J. Am. Chem. Soc., 1999, 121, 11432–11447 CrossRef.
- J. Y. Yang, R. M. Bullock, W. J. Shaw, B. Twamley, K. Fraze, M. R. DuBois and D. L. DuBois, Mechanistic Insights into Catalytic H2 Oxidation by Ni Complexes Containing a Diphosphine Ligand with a Positioned Amine Base, J. Am. Chem. Soc., 2009, 131, 5935–5945 CrossRef PubMed.
- A. D. Wilson, R. K. Shoemaker, A. Miedaner, J. T. Muckerman, D. L. DuBois and M. R. DuBois, Nature of hydrogen interactions with Ni(II) complexes containing cyclic phosphine ligands with pendant nitrogen bases, Proc. Natl. Acad. Sci. U. S. A., 2007, 104, 6951–6956 CrossRef PubMed.
- D. Schröder, S. Shaik and H. Schwarz, Two-State Reactivity as a New Concept in Organometallic Chemistry, Acc. Chem. Res., 2000, 33, 139–145 CrossRef PubMed.
- R. Poli and J. N. Harvey, Spin forbidden chemical reactions of transition metal compounds. New ideas and new computational challenges, Chem. Soc. Rev., 2003, 32, 1–8 RSC.
- B. Yang, L. Gagliardi and D. G. Truhlar, Transition states of spin-forbidden reactions, Phys. Chem. Chem. Phys., 2018, 20, 4129–4136 RSC.
- M. Swart and M. Gruden, Spinning around in Transition-Metal Chemistry, Acc. Chem. Res., 2016, 49, 2690–2697 CrossRef CAS PubMed.
- R. L. Yson, J. L. Gilgor, B. A. Guberman and S. A. Varganov, Protein induced singlet–triplet quasidegeneracy in the active site of [NiFe]-hydrogenase, Chem. Phys. Lett., 2013, 577, 138–141 CrossRef CAS.
- D. S. Kaliakin, R. R. Zaari and S. A. Varganov, Effect of H2 Binding on the Nonadiabatic Transition Probability between Singlet and
Triplet States of the [NiFe]-Hydrogenase Active Site, J. Phys. Chem. A, 2015, 119, 1066–1073 CrossRef CAS.
- A. D. Becke, Density-functional exchange-energy approximation with correct asymptotic behavior, Phys. Rev. A, 1988, 38, 3098–3100 CrossRef CAS.
- 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.
- D. Andrae, U. Häußermann, M. Dolg, H. Stoll and H. Preuß, Energy-adjustedab initio pseudopotentials for the second and third row transition elements, Theor. Chim. Acta, 1990, 77, 123–141 CrossRef CAS.
- R. Seeger and J. A. Pople, Self-consistent molecular orbital methods. XVIII. Constraints and stability in Hartree–Fock theory, J. Chem. Phys., 1977, 66, 3045–3050 CrossRef CAS.
- S. Chen, S. Raugei, R. Rousseau, M. Dupuis and R. M. Bullock, Homogeneous Ni Catalysts for H2 Oxidation and Production: An Assessment of Theoretical Methods, from Density Functional Theory to Post Hartree−Fock Correlated Wave-Function Theory, J. Phys. Chem. A, 2010, 114, 12716–12724 CrossRef CAS.
- S. Maeda, K. Ohno and K. Morokuma, Systematic exploration of the mechanism of chemical reactions: the global reaction route mapping (GRRM) strategy using the ADDF and AFIR methods, Phys. Chem. Chem. Phys., 2013, 15, 3683–3701 RSC.
- S. Maeda, T. Taketsugu, K. Morokuma and K. Ohno, Anharmonic Downward Distortion Following for Automated Exploration of Quantum Chemical Potential Energy Surfaces, Bull. Chem. Soc. Jpn., 2014, 87, 1315–1334 CrossRef.
- M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. J. A. Montgomery, J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian, Inc., Wallingford CT, 2016.
- A. V. Marenich, C. J. Cramer and D. G. Truhlar, Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions, J. Phys. Chem. B, 2009, 113, 6378–6396 CrossRef CAS.
- S. Chiodo and N. Russo, Determination of spin-orbit coupling contributions in the framework of density functional theory, J. Comput. Chem., 2008, 29, 912–920 CrossRef CAS.
- S. G. Chiodo and M. Leopoldini, MolSOC: A spin–orbit coupling code, Comput. Phys. Commun., 2014, 185, 676–683 CrossRef CAS.
- M.-H. Ho, R. Rousseau, J. A. S. Roberts, E. S. Wiedner, M. Dupuis, D. L. DuBois, R. M. Bullock and S. Raugei, Ab Initio-Based Kinetic Modeling for the Design of Molecular Catalysts: The Case of H2 Production Electrocatalysts, ACS Catal., 2015, 5, 5436–5452 CrossRef CAS.
- E. S. Rountree and J. L. Dempsey, Potential-Dependent Electrocatalytic Pathways: Controlling Reactivity with pKa for Mechanistic Investigation of a Nickel-Based Hydrogen Evolution Catalyst, J. Am. Chem. Soc., 2015, 137, 13371–13380 CrossRef CAS.
- E. S. Wiedner, H. J. S. Brown and M. L. Helm, Kinetic Analysis of Competitive Electrocatalytic Pathways: New Insights into Hydrogen Production with Nickel Electrocatalysts, J. Am. Chem. Soc., 2016, 138, 604–616 CrossRef CAS PubMed.
- E. S. Wiedner, J. A. S. Roberts, W. G. Dougherty, W. S. Kassel, D. L. DuBois and R. M. Bullock, Synthesis and Electrochemical Studies of Cobalt(III) Monohydride Complexes Containing Pendant Amines, Inorg. Chem., 2013, 52, 9975–9988 CrossRef CAS PubMed.
- E. S. Wiedner, A. M. Appel, D. L. DuBois and R. M. Bullock, Thermochemical and Mechanistic Studies of Electrocatalytic Hydrogen Production by Cobalt Complexes Containing Pendant Amines, Inorg. Chem., 2013, 52, 14391–14403 CrossRef CAS.
- A. Miedaner, R. C. Haltiwanger and D. L. DuBois, Relationship between the bite size of diphosphine ligands and tetrahedral distortions of “square-planar” nickel(II) complexes: stabilization of nickel(I) and palladium(I) complexes using diphosphine ligands with large bites, Inorg. Chem., 1991, 30, 417–427 CrossRef CAS.
- M. R. Nimlos, C. H. Chang, C. J. Curtis, A. Miedaner, H. M. Pilath and D. L. DuBois, Calculated Hydride Donor Abilities of Five-Coordinate Transition Metal Hydrides [HM(diphosphine)2]+ (M = Ni, Pd, Pt) as a Function of the Bite Angle and Twist Angle of Diphosphine Ligands, Organometallics, 2008, 27, 2715–2722 CrossRef CAS.
- D. D. M. Wayner and V. D. Parker, Bond energies in solution from electrode potentials and thermochemical cycles. A simplified and general approach, Acc. Chem. Res., 1993, 26, 287–294 CrossRef CAS.
Footnotes |
† Present address: Center for Computational Sciences, University of Tsukuba, 1-1-1 Tennodai, Tsukuba, Ibaraki 305-8577, Japan. |
‡ Present address: Department of Chemical System Engineering, The University of Tokyo, Tokyo 113-8656, Japan. |
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.