DOI:
10.1039/D5QO01184A
(Research Article)
Org. Chem. Front., 2025, Advance Article
Bis-guanidinate magnesium amide-catalyzed hydroboration of nitriles and isocyanides via a substrate-assisted pathway
Received
16th August 2025
, Accepted 23rd October 2025
First published on 24th October 2025
Abstract
The bis-guanidinate amido magnesium(II) compound [LMgN(SiMe3)2; L = {(ArHN)(ArN)–C
N–C = (NAr)(NHAr)}; [Ar = 2,6-Et2-C6H3] (Mg-1) has been employed as a pre-catalyst for the dihydroboration of nitriles, and as a catalyst for the dihydroboration of isocyanides, affording N,N-bis(boryl)amines and N,C-di(boryl)amines in good yields. The protocol introduces the first examples of nitrile and isocyanide reduction via a substrate-assisted pathway. The active catalyst for nitrile hydroboration, LMg-N(SiMe3)2·3-Me-C6H4CN (Mg-2), and the intermediate for isocyanide reduction, LMg-N(SiMe3)2·2,6-Me2-C6H3NC (Mg-3), were isolated and thoroughly characterized by multinuclear NMR spectroscopy and single-crystal X-ray diffraction. Based on these findings, plausible mechanistic cycles for both nitrile and isocyanide dihydroboration reactions have been proposed.
Introduction
The reduction of nitriles (R–C
N) and isocyanides (R–N
C) plays a pivotal role in synthetic chemistry.1–3 This process yields primary amines from nitriles and secondary amines from isocyanides, both of which are highly valuable due to their extensive applications. These compounds serve as essential feedstocks for agrochemicals, pharmaceuticals, dyes, and drug production, making them vital to industry and a key focus for synthetic chemists.4–13
Traditional ways for this kind of reduction include treatment of nitriles with metal hydrides,14 which may involve potential drawbacks such as use of pyrophoric and expensive reagents, poor selectivity,15,16 and direct hydrogenation with combustible hydrogen gas,17 which requires harsh reaction conditions18 often concerning about safety issues. In this context, metal-catalyzed hydroboration has proved to be more advantageous over traditional stoichiometric synthetic procedures by producing higher selectivity, milder reaction conditions, and reduced waste generation, and has emerged as a new alternative towards the reduction of various unsaturated organic moieties.13,19–24 Metal-catalyzed B–H or Si–H additions to the unsaturated organic functionalities are highly dominated by molecular metal hydrides following a hydride pathway;14,25 however, a few examples of hydride-free pathway are also reported.26–28 Despite extensive studies on transition,1,13,29–48 main group,1,48–65 and rare earth metal catalyzed1,48,66 hydroboration of nitriles and isocyanides, reports on magnesium-catalyzed hydroboration of the same, which remained scarce.56,57,63
Magnesium, being eco-friendly, non-toxic, earth-abundant, and biocompatible,67–71 offers significant potential in this domain. However, despite numerous advancements in the catalytic as well as stoichiometric applications of magnesium-based complexes,1,21,28,72,73 only three studies have explored its catalytic application in hydroboration of nitriles57 and isocyanides,63 underscoring a critical gap in this field.
The first report on metal-catalyzed nitrile dihydroboration was documented by the Nikonov group in 2012 using a molybdenum-based catalyst, (2,6-iPr2-C6H3N)Mo(H)(Cl)(PMe3)3.43 In 2016, the Hill group introduced a β-diketiminate (Nacnac) magnesium complex, [CH{C(Me)NAr}2Mg-nBu] (Ar = 2,6-iPr2C6H3), as a pre-catalyst for nitrile hydroboration, representing the first example with magnesium (Fig. 1). Notably, the same group reported the first metal-catalyzed hydroboration of isocyanides in 2015 using the same pre-catalyst, with both cases involving in situ generation of magnesium hydride species as the active catalyst. In 2018, Ma and coworkers reported unsymmetrical-diketiminate-stabilized Mg(I) dimer catalyzed double hydroboration of nitriles, though they couldn't mention any mechanistic insights on the catalytic transformation.56
 |
| | Fig. 1 Magnesium-catalyzed dihydroboration of nitriles and isocyanides. | |
Recently, the Trovitch group reported cobalt-catalyzed nitriles dihydroboration,38 in which authors introduced a chelate-assisted hydroboration pathway. To our knowledge, there have been no reports on the magnesium catalyzed hydroboration of nitriles and isocyanides involving chelate- assisted or substrate-driven pathway. This scarcity of studies on magnesium-catalyzed hydroboration of nitriles and isocyanides via substrate-assisted pathway, coupled with limited mechanistic insights, underscores the need to design innovative magnesium complexes and explore alternative pathways to unlock the full potential of this catalytic transformation.
In this context, we employed bis-guanidinate Mg(II) amide [LMgN(SiMe3)2] (Mg-1)74 as a pre-catalyst for the hydroboration of the nitriles and a catalyst for the dihydroboration of isocyanides with good to excellent yields. We have isolated and characterized the active species LMg-N(SiMe3)2·3-Me-C6H4CN (Mg-2), and LMg-N(SiMe3)2·2,6-Me2-C6H3NC (Mg-3) with various spectroscopic techniques (1H, 13C{1H}, 29Si{1H}) and XRD analysis. Moreover, based on the above findings, we established the first example of amido magnesium complex-catalyzed hydroboration of nitriles and isocyanides, with a mechanistic insight into the substrate-assisted pathway.
Results and discussion
Dihydroboration of nitriles
Our group previously developed a “Nacnac” analogue – specifically, a tetra-N-aryl substituted bis-guanidine ligand and its coordination chemistry.64,75,76 Recently, we synthesized bis-guanidinate magnesium amide complex (Mg-1) and employed it as a catalyst for the C–C coupling of terminal alkynes with carbodiimide (CDI) (Fig. 2).74
 |
| | Fig. 2 Bis-guanidinate stabilized Mg(II) amide (Mg-1). | |
Inspired by the work of the Hill group on magnesium-catalyzed dihydroboration of nitriles,57 we investigated the use of Mg-1 as a pre-catalyst for nitrile dihydroboration. Our initial experiment began with m-tolunitrile (1c) as a model substrate and pinacolborane (HBpin) as a hydroborating agent, to optimize the catalytic transformation. A control reaction without a catalyst, performed at 80 °C for 24 h under neat conditions, showed no conversion. However, under identical conditions, the presence of 5 mol% of Mg-1 afforded quantitative conversion to the corresponding N,N-bis(boryl)amine (2c), confirmed by 1H NMR spectroscopy (Table 1, entry 2). Reducing the catalyst loading to 1 mol% under neat conditions at 60 °C achieved full conversion within 12 h (Table 1, entry 4). Further decrease in the reaction time led to reduced NMR conversions (Table 1, entries 5 and 6).
Table 1 Optimization table of magnesium catalyzed dihydroboration of m-tolunitrilea

|
| Entry |
Cat. |
mol% |
Solvent |
Temp. (°C) |
Time (h) |
Conv.b (%) |
| Reaction conditions: m-tolunitrile (0.1 mmol, 1.0 equiv.), pinacolborane (HBpin) (0.2 mmol, 2.0 equiv.), catalyst Mg-1, Mg-2 and Mg[N(SiMe3)2]2 (x mol%). Conversions for the reduction of m-tolunitrile to N,N′-bis-borylated amine were examined by 1H and 13C{1H} NMR spectroscopy based on the consumption of starting material and formation of characteristic new proton resonance for {CH2N(Bpin)2} moiety. |
| 1 |
— |
— |
Neat |
80 |
24 |
— |
| 2 |
Mg-1 |
5 |
Neat |
80 |
24 |
>99 |
| 3 |
Mg-1 |
3 |
Neat |
70 |
12 |
>99 |
| 4 |
Mg-1 |
1 |
Neat |
60 |
12 |
>99 |
| 5 |
Mg-1 |
1 |
Neat |
60 |
11 |
94 |
| 6 |
Mg-1 |
1 |
Neat |
60 |
6 |
55 |
| 7 |
Mg-1 |
1 |
Neat |
rt |
24 |
30 |
| 8 |
Mg-1 |
0.5 |
Neat |
60 |
12 |
75 |
| 9 |
Mg-1 |
0.5 |
Neat |
60 |
24 |
75 |
| 10 |
Mg-1 |
0.5 |
Neat |
80 |
24 |
75 |
| 11 |
Mg-1 |
1 |
C6D6 |
60 |
14 |
>99 |
| 12 |
Mg-2 |
1 |
C6D6 |
60 |
10 |
>99 |
| 13 |
Mg[N(SiMe3)2]2 |
1 |
Neat |
60 |
12 |
10 |
The same reaction produced quantitative conversion upon heating at 60 °C for 14 h in C6D6 (Table 1, entry 11). Additionally, we tested the activity of Mg-2 (vide infra) for the dihydroboration reaction, which produced quantitative conversion at 80 °C, within 10 h in C6D6 (Table 1, entry 12). We also tested the catalytic activity of magnesium bis(trimethylsilyl)amide [Mg{N(SiMe3)2}2]77 reagent, which showed only 10% conversion to the diborylated product (Table 1, entry 13), highlighting the superior activity of a heteroleptic molecular magnesium complex like Mg-1, over a homoleptic magnesium complex [Mg{N(SiMe3)2}2]. Further reduction in catalyst loading led to decreased conversions, even at elevated temperature and prolonged reaction time (Table 1, entries 7–10).
With optimized reaction conditions in hand, we explored the substrate scope. A variety of nitriles (1a–1p), including electron-donating (1a–1g), electron-withdrawing (1h–1n) substituted aryl nitriles, as well as alkyl nitriles (1o, 1p), were successfully dihydroborated with HBpin in the presence of Mg-1. The corresponding N,N-bis(boryl)amines (2a–2p) were obtained with quantitative conversion (Table 2) except for 2p (88%). Nitrile 1a underwent complete conversion to 2a within 6 h. To probe substituent effects, we examined 1c (electron-donating) and 1m (electron-withdrawing) under shorter reaction times (<12 h), which resulted in comparatively lower conversions (11 h, ∼90%). Based on these observations, subsequent substrates were screened under standardized overnight heating (up to 12 h), ensuring quantitative conversions. Furthermore, to assess the effect of electronic substitution on the initial rate, we monitored the hydroboration of 1c and 1m at 3 h under optimal conditions in C6D6 (see SI, Fig. S20 and S42). Both substrates displayed nearly identical NMR conversions (32–34%), indicating comparable initial reaction rates. These observations suggest that electronic effects of the substituents have little influence on the reaction rate, and the reduced rates in the substituted nitriles are instead attributed to steric hindrance during catalyst–substrate complex formation (see stoichiometric reactions). The aliphatic nitrile 1o gave >99% NMR conversion, while its aryl analogue 1p gave a slightly lower yield (88%). All products were characterized by 1H and 13C{1H} NMR spectroscopy. New signals in the range of 2.86–4.38 ppm in the 1H NMR spectra, corresponding to the CH2N(Bpin)2 moiety, confirmed successful conversion. The yield of 2p was determined using mesitylene as an internal standard [see SI, Fig. S47 and S48].
Table 2 Substrate scope for Mg-1 catalyzed dihydroboration of nitrilesa,b,c
| Reaction conditions: nitriles (0.2 mmol, 1.0 equiv.), pinacolborane (HBpin) (0.4 mmol, 2.0 equiv.), catalyst Mg-1 (1 mol% %), heating at 60 °C, 12 h (6 h for 1a) stirring under neat conditions. Conversion for the product was examined by 1H and 13C{1H} NMR spectroscopy based on the consumption of starting material and formation of characteristic new proton resonance for the (CH2N(Bpin)2) moiety of products 2a–2p. For compounds 2i and 2p, mesitylene is used as an internal standard. |
 |
Dihydroboration of isocyanides
Encouraged by our success with nitriles and noting the limited reports on isocyanide hydroboration,62,63 we evaluated the catalytic activity of Mg-1 for the dihydroboration of isocyanide. Using 2,6-dimethylphenyl isocyanide (3a) as a model substrate, a catalyst-free reaction with HBpin at 80 °C for 24 h in benzene-d6 yielded no conversion (Table 3, entry 1). In contrast, using 10 mol% Mg-1 resulted in quantitative formation of N,C-di(boryl)amine (4a) (Table 3, entry 2), as confirmed by 1H NMR spectroscopy. Decreasing the catalyst loading to 6 mol% still provided full conversion at 70 °C within 12 h (Table 3, entry 4). The same reaction showed no change in the NMR conversion under neat conditions (Table 3, entry 5). Further reduction in the catalyst load decreased the conversion significantly (Table 3, entry 6). The reagent Mg[N(SiMe3)2]277 produced only 30% conversion under standard reaction conditions (Table 3, entry 7).
Table 3 Optimization table of magnesium catalyzed hydroboration of isocyanidea,b,c

|
| Entry |
Cat |
mol% |
Solvent |
Temp. (°C) |
Time (h) |
Conv.b (%) |
| Optimization reaction conditions: 2,6-dimethylphenyl isocyanide (3a) (0.1 mmol, 1.0 equiv.), pinacolborane (HBpin) (0.21 mmol, 2.1 equiv.), catalyst Mg-1 or Mg[N(SiMe3)2]2 (x mol%). Conversion of 3a to the corresponding 1,2-diborylated amine (4a) was examined by 1H and 13C{1H} NMR spectroscopy based on the consumption of starting material and formation of characteristic new proton resonance for the {CH2N(Bpin)2} moiety. Reactions were performed with 3b and 3c (0.1 mmol, 1.0 equiv.), pinacolborane (HBpin) (0.21 mmol, 2.1 equiv.), catalyst Mg-1 (6 mol%), heated at 70 °C, for 12 h. Conversions to 1,2-diborylated amines (4b, 4c)were examined by 1H and 13C{1H} NMR spectroscopy based on the formation of a characteristic new proton resonance for {CH2(Bpin)N(Bpin)} moiety. |
| 1 |
— |
— |
C6D6 |
80 |
24 |
— |
| 2 |
Mg-1 |
10 |
C6D6 |
80 |
24 |
>99 |
| 3 |
Mg-1 |
7 |
C6D6 |
70 |
12 |
>99 |
| 4 |
Mg-1 |
6 |
C6D6 |
70 |
12 |
>99 |
| 5 |
Mg-1 |
6 |
Neat |
70 |
12 |
>99 |
| 6 |
Mg-1 |
5 |
C6D6 |
70 |
24 |
80 |
| 7 |
Mg[N(SiMe3)2]2 |
6 |
C6D6 |
70 |
12 |
30 |
 |
Next, we investigated other isocyanides: 2-naphthyl isocyanide (3b) showed 70% conversion to 4b, while cyclohexyl isocyanide (3c) underwent complete conversion to 4c (Table 3). These products were confirmed by 1H and 13C{1H} NMR spectroscopy. New singlets in the 2.74–3.61 ppm range, attributed to the N(Bpin)CH2Bpin moiety, confirmed successful formation (see SI). We attempted dihydroboration of isocyanide substrates 3d, 3e, and 3f; however, the corresponding borylated products 4d, 4e, and 4f were obtained in low conversions (SI, Table S1).
Stoichiometric reactions for the hydroboration of nitriles and isocyanides
To gain mechanistic insights into the bis-guanidinate magnesium amide-catalyzed hydroboration reactions of nitriles and isocyanides, we executed a series of control experiments. Previously, the Hill group reported β-diketiminate (Nacnac) stabilized magnesium alkyl as a pre-catalyst for the hydroboration of organic nitriles,57 involving the in situ formation of magnesium hydride. It is worthy to note, that molecular magnesium hydrides can be accessed by the reaction between magnesium amide and silane reagents,14,78–84 however, a 1
:
3 stoichiometric reaction of Mg-1 and HBpin at 90 °C for 24 h indicated the formation of the Mg1-HBpin adduct in the 1H NMR spectrum (see SI, Fig. S7) (Scheme 1i). In addition, we observed no conversion of HBpin into BH3 in the presence of Mg-1 (see SI, Fig. S13) which rules out the any hidden catalysis85,86 for the hydroboration reactions. A 1
:
1 stoichiometric reaction between Mg-1 and m-tolunitrile (1c) in C6D6 produced the monomeric nitrile coordinated adduct LMg-N(SiMe3)2·3-Me-C6H4CN Mg-2 within 5 min of the reaction at room temperature (Scheme 1ii). The compound Mg-2 was isolated and characterized by multinuclear NMR spectroscopy and single-crystal X-ray diffraction analysis. The 1H NMR spectrum of the compound Mg-2 displayed a downfield shift of the methyl groups of the N(SiMe3)2 moiety at 0.35 ppm than its parent precursor Mg-1 at 0.01 ppm in C6D6,74 indicating the coordination of the electron-withdrawing nitrile group. A considerable downfield shift has been observed for the Ar–NH moiety of the ligand backbone at 5.03 ppm (Mg-1; Ar–NH: 1H NMR δ 4.89 ppm) in the 1H NMR spectrum, indicating the Mg-2 formation. 13C{1H} NMR spectrum also displayed a significant shift of the N(SiMe3)2 carbons at 5.7 ppm than the Mg-1 at 4.4 ppm in C6D6 [see SI, Fig. S2]. Additionally, one singlet signal at −7.93 ppm in the 29Si{1H} NMR spectrum shows a slight shift from its parent precursor at −7.22 ppm in C6D6, indicating the formation of the new species Mg-2 [see SI, Fig. S1, S2, and S3].
 |
| | Scheme 1 Control experiments for dihydroboration of nitrile. | |
Though some isocyanide adducts of ligated magnesium complexes are reported in the literature,87 to our knowledge, Mg-2 is the first example of a nitrile adduct of a bis-guanidinate-stabilized magnesium amide complex. The single crystals suitable for X-ray diffraction analysis were grown inside a J. Young valve NMR tube in C6D6 at room temperature. The X-ray diffraction study shows that the compound crystallizes in a monoclinic system with a P21/n space group. The molecular structure shows a monomeric unit with a magnesium centre surrounded by four N atoms: two from the chelated bis-guanidinate ligand, and others from the N(SiMe3)2 and the nitrile (CN) group, attaining a distorted tetrahedral geometry (Fig. 3). The Mg1–N6 bond distance of the compound Mg-2 is 2.0036 (14) Å, significantly longer than the Mg–N(SiMe3)2 bond distance of 1.9503 (14) Å in Mg-1,74 which is due to the enhanced steric hindrance at the magnesium centre in Mg-2 than Mg-1.
 |
| | Fig. 3 Molecular structure of Mg-2. The thermal ellipsoids are shown at probability 50%, and H atoms except H4 and H5, and the ethyl substitutions on the N-aryl groups of the bis-guanidinate ligand are omitted for clarity. The selected bond lengths (Å) and bond angles (°): Mg1–N6 2.0036 (14), Mg1–N2 2.0427 (13), Mg1–N1 2.0447 (13), Mg1–N7 2.1546 (14), Si1–N6 1.7014 (13), Si2–N6 1.7039 (13), N7–C3 1.140 (2), N1–Mg1–N2 89.64 (5), N1–Mg1–N6 119.13 (6), N2–Mg1–N6 134.40 (6), Mg1–N7–C3 156.68 (14), Si1–N6–Mg1 120.66 (7), Si2–N6–Mg1 117.55 (7), Si1–N6–Si2 121.56 (8). | |
The N atom of the nitrile is attached to the magnesium through a coordinate bond, which is evident from the longer Mg1–N7 bond distance of 2.1546 (14) Å than the purely covalent Mg1–N6 bond distance of 2.0036 (14) Å. The C3–N7 bond distance is 1.140 (2) Å, which is well in agreement with the C
N moiety of the nitrile functional group, which confirms that the nitrile group is intact and coordinated to the magnesium centre. The N1–Mg1–N2 bite angle in the Mg-2 is 89.64 (5)°, which is acute than the bite angle of both the bis-guanidinate [94.36 (6)°]74 as well as Nacnac stabilized amido magnesium complexes [95.11(6)°].88
Attempts to isolate N-borylimine intermediates from catalytic reactions of 1c with HBpin were unsuccessful, likely due to their high reactivity than nitriles, leading to full reduction of N,N-bis(boryl) amine.89,90 Notably, there have been a few reports on monohydroboration of nitriles.44,91–96
Surprisingly, in a 1
:
1 stoichiometric reaction of Mg-2 and HBpin at 60 °C in C6D6 for 30 min, we observed the formation of an entirely new magnesium N-boryl imine compound Int-1 (Scheme 1iii), where a N-borylated imine part is stabilized by the guanidinate magnesium amide counterpart. The Int-1 was obtained with 50% NMR yield and was identified through 1H NMR spectroscopy [see SI, Fig. S9]. A diagnostic singlet signal at 8.72 ppm in the 1H NMR spectrum of Int-1, in C6D6, indicates the Mg-N(Bpin)
C–H moiety of the newly formed borylated imine. This is comparable to the Hill group's reported Nacnac stabilized magnesium aldimidoborohydride, where the imine N
C–H proton resonates near 9.27 ppm in the 1H NMR spectrum.57 Further heating of the reaction mixture started to produce the fully reduced N,N-bis(boryl) amine. A very careful addition of m-tolunitrile (1c) to the mixture of Mg-2 and Int-1, at room temperature slowly formed the N-borylated imine (2c′) and regenerated Mg-2, within 12 h (Scheme 1iii) [see SI, Fig. S11] with a quantitative conversion, which implies that the Mg-2 and Int-1 are in equilibrium in the reaction mixture. Further isolation of the 2c′ was unsuccessful due to the extreme reactivity of the compound; it couldn't be stabilized in its imine form. Despite several attempts, we were unable to isolate the Int-1 in its solid-state form. All the sets of stoichiometric reactions are summarized in the stacked 1H NMR representation in Fig. 4.
 |
| | Fig. 4 Stacked 1H NMR spectra for the stoichiometric reactions. | |
The overall study suggests that the reaction proceeds via an in situ formation of an active magnesium hydride species by Mg–N/H–B metathesis upon treatment of HBpin with Mg-2 (Scheme 3, vide infra). The formation of the adduct is a crucial step, as it weakens the Mg–N(SiMe3)2 bond [Mg-2; Mg1–N6: 2.0036(14) Å, Mg-1; Mg–N: 1.9503(14) Å], making it more susceptible to metathesis with HBpin. This process drives the overall reaction forward, indicating a substrate-assisted pathway.
Next, a 1
:
2 stoichiometric reaction of Mg-2 and HBpin in the presence of m-tolunitrile at 60 °C in C6D6 formed the N,N-bis(boryl) amine 2c, and along with the regeneration of the compound Mg-2 (Scheme 1iv) within 12 h [see SI, Fig. S10]. The formation of the compound 2c was confirmed by 1H NMR spectroscopy. The disappearance of the singlet signal at 8.72 ppm in the 1H NMR spectrum in C6D6 and the formation of a new signal at 4.63 ppm indicate the CH2 protons of the N(Bpin)2-CH2-Ar moiety, which confirms the production of the compound 2c.
A 1
:
1 stoichiometric reaction between Mg-1 and 2,6-dimethylphenyl isocyanide (3a) in C6D6 readily formed the monomeric isocyanide adduct LMg-N(SiMe3)2·2,6-Me2-C6H3NC Mg-3 within 5 min at room temperature (Scheme 2i).
 |
| | Scheme 2 Control experiments for dihydroboration of isocyanide. | |
The compound Mg-3 was isolated and characterized by multinuclear NMR spectroscopy and single-crystal X-ray diffraction. Mg-3 has shown a significant downfield shift of the 18H of the N(SiMe3)2 group at 0.18 ppm (Mg-1; N(SiMe3)2: δ 0.01 ppm) in the 1H NMR spectrum in C6D6, indicating the coordination of the electron-withdrawing N
C group. A singlet signal at 2.00 ppm, in the 1H NMR spectrum, corresponding to the methyl groups of the 2,6-dimethylphenyl isocyanide. Additionally, the 29Si{1H} NMR spectrum displays a singlet signal at −7.69 ppm in C6D6, significantly shifted from its parent precursor Mg-1 at −7.22 ppm, confirming the formation of the Mg-3 [see SI, Fig. S4–S6]. Diffractable single crystals of the compound Mg-3 were grown inside the J. Young valve NMR tube at room temperature in C6D6. The X-ray analysis shows the compound Mg-3 crystallizes in a monoclinic system with a P-1 space group. The molecular structure displayed a monomeric structure where one molecule of 2,6-dimethylphenyl isocyanide is coordinated to the magnesium centre of Mg-1, through the carbon atom, making it a four-coordinated magnesium complex with a distorted tetrahedral geometry (Fig. 5). To the best of our knowledge, Mg-3 is the first example of an isocyanide adduct of a bis-guanidinate magnesium complex. The Mg1–N6 bond distance is 2.0077 (13) Å, significantly longer than the parent precursor Mg-1 (Mg-N(SiMe3)2; 1.9503 (14) Å), corresponding to the enhanced steric hindrance due to the increased coordination number to the magnesium centre. The C1–N7 bond distance in the coordinated isocyanide counterpart is 1.159 (2) Å, slightly longer than the C–N bond distance [N3–C38 1.145 (3) Å] in the isocyanide adduct of the analogous Nacnac stabilized magnesium siloxide,97 reported by Hill and coworkers. The C1–N7–C2 bond angle is 178.53 (16)°, which implies a close to linear framework of the isocyanide moiety. The longer C1–N7 bond distance indicates the ease of the B–H addition to the isocyanide R–N
C: bond. Similar to the Mg-2, the N1–Mg1–N2 bite angle in the compound Mg-3 is 90.53 (5)°, which is acute than the bite angles of both the Mg-174 as well as the analogous Nacnac magnesium amide complex (vide supra).88
 |
| | Fig. 5 Molecular structure of compound Mg-3. The thermal ellipsoids are shown at probability 50%, and H atoms except H4 and H5 and the ethyl substitutions on the N-aryl groups of the bis-guanidinate ligand are omitted for clarity. The selected bond lengths (Å) and bond angles (°): Mg1–N6 2.0077 (13), Mg1–N2 2.0492 (12), Mg1–N1 2.0681 (12), Mg1–C1 2.3599 (16), Si1–N6 1.7067 (13), Si2–N6 1.7132 (13), N7–C1 1.159 (2), N1–Mg1–N2 90.53 (5), N1–Mg1–N6 132.50 (5), N2–Mg1–N6 120.99 (5), Mg1–C1–N1 176.39 (12), Si1–N6–Mg1 120.17 (7), Si2–N6–Mg1 120.89 (7), Si1–N6–Si2 118.88 (7). | |
A subsequent stoichiometric reaction of compound Mg-3 with 2 equiv. of HBpin at 70 °C for 12 h formed the N,C-di(boryl)amine 4a and Mg-1 with a quantitative conversion (Scheme 2ii), verified through 1H NMR spectroscopy [see SI, Fig. S12].
Kinetic experiments and discussions
To justify the mechanistic implications of the stoichiometric outcomes, a kinetic study via comparison of concentration profiles98 for the Mg-1-catalyzed nitrile's dihydroboration was conducted. The m-tolunitrile(1c) was chosen as a standard substrate because of its indicative signals for its consumption and the formation of the product in the 1H NMR spectrum. All the reactions were carried out at 343 K and were monitored by 1H NMR spectroscopy up to 80% conversion with respect to mesitylene used as an internal standard, using 10 mol% of the Mg-1 with various ratios of m-tolunitrile to HBpin.
A 1
:
2 stoichiometric ratio of m-tolunitrile (1c) (0.16 M) and HBpin (0.33 M), with varying catalyst load of Mg-1 adjusted at 10 and 7 mol%, apparently displayed a first order kinetics, (Fig. 6A) consistent with other Nacnac stabilized alkaline earth metal complexes mediated catalysis reactions.57,99–101 This observation indicates involvement of a single isolated magnesium center, which is involved in the rate determination of the reaction.
 |
| | Fig. 6 (A) ln[m-tolunitrileConcn] vs. Time plot, at different loadings of Mg-1. (B) [Product Concn] vs. Time plot, at pseudo-first order condition with respect to m-tolunitrile. | |
A series of pseudo-first-order reactions was conducted using a 20-fold excess of m-tolunitrile (3.33 M) and a catalyst loading of 0.0167 M while varying HBpin concentrations from 0.33 M to 1.33 M. The reaction rate displayed a clear first-order kinetics at lower HBpin concentrations (0.33 M and 0.66 M) but deviated to a near-zero-order dependence at higher concentrations (1.00 M, 1.16 M, and 1.33 M) (Fig. 6B). It was observed that the rate of product formation increased significantly from 0.33 M to 0.66 M HBpin, whereas only a marginal change was observed between 1.00 M and 1.33 M, indicating saturation kinetics, consistent with the formation of a catalyst–substrate intermediate species (designated as Int-1) similar to the Hill group's reported magnesium-catalyzed nitriles dihydroboration.57 This saturation behavior suggests a Michaelis–Menten–type102,103 kinetic profile, where the reaction of Mg-2 and HBpin proceeds through the formation of an active catalyst–HBpin intermediate (ca. Int-1), which precedes the rate-determining step.
| Catalyst + HBpin ⇌ [Catalyst-HBpin] → Product |
At lower concentrations of HBpin (0.33 M, 0.67 M), the active catalyst Mg-2 remains unsaturated, and the rate increases proportionally with HBpin concentration, indicating a first-order kinetics arising from the rate-limiting formation of Int-1; however, at higher concentrations of HBpin (1.00 M to 1.33 M), catalyst saturation occurs, and the rate becomes independent of the concentration of HBpin, showing zero-order kinetics. This transition highlights the key role of catalyst–substrate complexation in governing the observed rate behavior. A stacked 1H NMR study (Fig. 7) of the reaction pseudo-first-order conditions with respect to nitrile (20 equiv.) revealed the formation of an N-boryl imine intermediate (2c′) at high nitrile concentrations. This observation aligns with outcomes from stoichiometric reactions. As the reaction proceeds and nitrile is consumed, 2c′ formation drops while the final product 2c starts to appear. The 2c′ is observed only under stoichiometric and kinetic conditions, not in catalytic runs, suggesting it is a short-lived intermediate.
 |
| | Fig. 7 Stacked 1H NMR spectra (400 MHz) for the reaction of m-tolunitrile (1c) (3.33 M, 20 equiv.), HBpin (0.33 M, 2 equiv.), and Mg-1 (10 mol%, 0.0167 M) in C6D6. Spectra were recorded at 70 °C, in different time intervals from 30 min to 360 min. 2c′; 3-CH3-C6H4-CH N(Bpin). 2c; 3-CH3-C6H4-CH2-N(Bpin)2. | |
To investigate the effect of nitrile concentration on the reaction rate, a series of catalytic reactions was conducted using 0.0167 M Mg-1 under pseudo-first-order conditions with respect to HBpin (3.33 M, 20-fold excess), while varying the concentration of m-tolunitrile (0.167–0.67 M).
At lower nitrile concentrations (0.167 and 0.33 M), a zero-order kinetics was observed, indicating that nitrile's concentration plays no direct role in the rate-determining step and is only involved in the formation of the catalytically active species, Mg-2. However, at higher nitrile concentration (0.67 M), the reaction rate decreased despite maintaining a zero-order kinetics, suggesting that excess nitrile concentration inhibits the reaction by saturation of the Mg-2 active site and impeding HBpin access (Fig. 8A).
 |
| | Fig. 8 (A) [Product Concn] vs. time plot, at pseudo-first order condition with respect to HBpin, linear fit. (B) ln[Nitrile Concn] vs. Time plot, for electron withdrawing and electron donating substituted aryl nitriles, linear fit. | |
A further study examined the effect of electron-donating and electron-withdrawing groups on the reaction mechanism (Fig. 8B). The dihydroboration of 2,4-difluorobenzonitrile (1i) proceeded with a slightly decreased rate with respect to m-tolunitrile (1c), following first-order kinetics. The slightly reduced reaction rate for 1i is likely due to its strong electron-withdrawing nature and more steric hindrance, which decreases the availability of the nitrile lone pair for coordination with Mg-1. This in turn, inhibits the formation of [Mg-1 nitrile] adduct, a key step in the overall reaction pathway. Further study of dihydroboration of both the electron-withdrawing (1m) and the electron-donating (1c) substituted nitriles within initial 3 h, exhibited nearly similar reaction rate, highlighting the importance of the formation [Mg-1 nitrile] adduct.
In summary, kinetic studies show that the reaction follows first-order kinetics with respect to the magnesium catalyst, indicating the involvement of a single active species governing the reaction rate. Nitriles exhibit zero-order kinetics; however, they inhibit the reaction rate at higher concentrations due to catalyst saturation. HBpin shows first-order kinetics at low concentrations, shifting to zero-order at higher concentrations, consistent with the formation of a catalyst–substrate intermediate, indicating a Michaelis–Menten-type behavior. Electron-withdrawing nitriles, such as 2,4-difluorobenzonitrile, suppress the reaction rate slightly by weakening coordination to the magnesium center. Overall, the catalytic cycle is primarily governed by the formation of Mg-2, and its reaction with HBpin.
The rate of the reaction follows eqn (1) and (2).
(At a lower concentration of HBpin)
| |
 | (1) |
(At a higher concentration of HBpin)
| |
 | (2) |
Catalytic cycle for the dihydroboration of nitriles
Based on stoichiometric reactions and the previous literature reports,38,104 we propose a (Scheme 3) a plausible mechanistic cycle for the Mg-1-catalyzed hydroboration of organic nitriles and isocyanides. As far as magnesium-catalyzed hydroboration is concerned, initially Mg-1 reacts with nitrile 1c to form the active species bis-guanidinate magnesium nitrile adduct Mg-2. The reaction of Mg-2 and HBpin, generates a short-lived magnesium imido intermediate Int-1′ via in situ hydride formation. This species subsequently undergoes rapid reaction with pinB-N(SiMe3)2 (Scheme 3) to afford magnesium amino-N-boryl imine intermediate (Int-1). The Int-1 simultaneously reacts with HBpin to form Int-2, which, upon reacting with m-tolunitrile (1c), produces the desired N,N′-bis-boryl amine (2c) and regenerates the active species Mg-2, hereby closing the catalytic cycle. Notably, at the outset, Int-1 remains in equilibrium with Mg-2, which, upon reacting with excess m-tolunitrile(1c), possibly produces N-boryl imine 2c′ along with reformation of Mg-2; however, this has not been observed in the catalytic reactions under optimal conditions. Unlike magnesium-catalyzed dihydroboration of nitriles, Mg-1 acts as an active catalyst for the dihydroboration of isocyanides. Initially, the Mg-1 reacts with 3a and forms intermediate Mg-3 via coordination of isocyanide to the magnesium centre, which reacts with 2 equiv. of HBpin to produce the 1,2-diborylated amine 4a, along with the regeneration of the active catalyst Mg-1, thereby closing the catalytic cycle (Scheme 3). The overall outcomes successfully established Mg-1 as a pre-catalyst for the nitriles dihydroboration; however, it acts as an active catalyst for the dihydroboration of isocyanides.
 |
| | Scheme 3 Mechanistic cycle of the Mg-1-catalyzed dihydroboration of nitrile and isocyanide. | |
Conclusion
In conclusion, we have successfully synthesized the first heteroleptic bis-guanidinate magnesium amide adducts bearing coordinated nitrile and isocyanide groups. Furthermore, we have demonstrated the magnesium amide-catalyzed hydroboration of nitriles and isocyanides via a substrate-assisted pathway. This catalytic protocol enables the formation of N,N-bis(boryl)amines and N,C-di(boryl)amines in good to excellent yields. Notably, we observed that in contrast to the homoleptic magnesium bis-amide, the heteroleptic N,N′-chelated bis-guanidinate magnesium amide acts as a superior catalyst for both the dihydroboration of nitriles and isocyanides. Comprehensive mechanistic investigations were established through various stoichiometric reactions, which led to the successful isolation of catalytically relevant intermediates and the active magnesium species, which were unambiguously identified using multinuclear NMR spectroscopy and single-crystal X-ray diffraction analysis. Notably, we were able to generate a highly reactive N-boryl imine intermediate in a stoichiometric reaction. These findings not only expand the scope of magnesium catalysis in bond-forming transformations but also provide critical mechanistic insights into substrate-assisted activation pathways in main-group catalysis.
Author contributions
S. M. carried out most of the synthesis and characterization, drafted the manuscript, and contributed to its subsequent and final versions. S. R. P. managed the supplementary data and assisted in catalyst preparation and kinetic studies. A. D. conducted the catalysis studies and contributed to manuscript drafting. S. N. conceptualized the project, acquired funding, supervised the work, and contributed to the final manuscript.
Conflicts of interest
The authors declare no conflict of interest.
Data availability
Supplementary information (SI): all the materials and instrumentation, and supplementary figures and tables. See DOI: https://doi.org/10.1039/d5qo01184a.
CCDC 2476567 and 2476571 contain the supplementary crystallographic data for this paper.105a,b
Acknowledgements
The authors thank the National Institute of Science Education and Research (NISER), Bhubaneswar, Odisha, Homi Bhabha National Institute (HBNI), Mumbai, Department of Atomic Energy (DAE), Govt. of India. The Science and Engineering Research Board (SERB), India (CRG/2021/007000) is acknowledged for providing financial support.
References
- D. Hayrapetyan and A. Y. Khalimon, Catalytic Nitrile Hydroboration: A Route to N,N-Diborylamines and Uses Thereof, Chem. – Asian J., 2020, 15, 2575–2587 CrossRef CAS PubMed.
- P. Patil, M. Ahmadian-Moghaddam and A. Dömling, Isocyanide 2.0, Green Chem., 2020, 22, 6902–6911 RSC.
- B. R. Barnett, C. E. Moore, A. L. Rheingold and J. S. Figueroa, Frustrated Lewis pair behavior of monomeric (boryl) iminomethanes accessed from isocyanide 1, 1-hydroboration, Chem. Commun., 2015, 51, 541–544 RSC.
- J. S. Carey, D. Laffan, C. Thomson and M. T. Williams, Analysis of the reactions used for the preparation of drug candidate molecules, Org. Biomol. Chem., 2006, 4, 2337–2347 RSC.
- G. P. Chiusoli and P. M. Maitlis, Metal-catalysis in industrial organic processes, Royal Society of Chemistry, 2019 Search PubMed.
- K. Hayes, Industrial processes for manufacturing amines, Appl. Catal., A, 2001, 221, 187–195 CrossRef CAS.
- J. Magano and J. R. Dunetz, Large-Scale Applications of Transition Metal-Catalyzed Couplings for the Synthesis of Pharmaceuticals, Chem. Rev., 2011, 111, 2177–2250 CrossRef CAS PubMed.
- R. Martin and S. L. Buchwald, Palladium-catalyzed Suzuki–Miyaura cross-coupling reactions employing dialkylbiaryl phosphine ligands, Acc. Chem. Res., 2008, 41, 1461–1473 CrossRef CAS PubMed.
- T. Renders, S. Van de Vyver and B. F. Sels, Bio-based amines through sustainable heterogeneous catalysis, Green Chem., 2017, 19, 5303–5331 RSC.
- C. Torborg and M. Beller, Catalysis, Recent applications of palladium–catalyzed coupling reactions in the pharmaceutical, agrochemical, and fine chemical industries, Adv. Synth. Catal., 2009, 351, 3027–3043 CrossRef CAS.
- N. Miyaura and A. Suzuki, Palladium-Catalyzed Cross-Coupling Reactions of Organoboron Compounds, Chem. Rev., 1995, 95, 2457–2483 CrossRef CAS.
- C. Pubill-Ulldemolins, A. Bonet, C. Bo, H. Gulyás and E. Fernández, A new context for palladium mediated B-addition reaction: an open door to consecutive functionalization, Org. Biomol. Chem., 2010, 8, 2667–2682 RSC.
- C. C. Chong and R. Kinjo, Catalytic Hydroboration of Carbonyl Derivatives, Imines, and Carbon Dioxide, ACS Catal., 2015, 5, 3238–3259 CrossRef CAS.
- M. M. Roy, A. A. Omana, A. S. Wilson, M. S. Hill, S. Aldridge and E. Rivard, Molecular main group metal hydrides, Chem. Rev., 2021, 121, 12784–12965 CrossRef CAS PubMed.
- Q. Wang, J. Guo and P. Chen, Complex transition metal hydrides for heterogeneous catalysis, Chem. Catal., 2023, 3, 100524 CAS.
- N. Klopčič, I. Grimmer, F. Winkler, M. Sartory and A. Trattner, A review on metal hydride materials for hydrogen storage, J. Energy Storage, 2023, 72, 108456 CrossRef.
- J. G. de Vries and C. J. Elsevier, Handbook of homogeneous hydrogenation, Weinheim: Wiley-VCH, 2007 Search PubMed.
- P. Andersson, Modern Reduction Methods, Wiley-VCH, 2008 Search PubMed.
- K. Kuciński and G. Hreczycho, Hydrosilylation and hydroboration in a sustainable manner: from Earth-abundant catalysts to catalyst-free solutions, Green Chem., 2020, 22, 5210–5224 RSC.
- M. L. Shegavi and S. K. Bose, Recent advances in the catalytic hydroboration of carbonyl compounds, Catal. Sci. Technol., 2019, 9, 3307–3336 RSC.
- M. Magre, M. Szewczyk and M. Rueping, s-Block Metal Catalysts for the Hydroboration of Unsaturated Bonds, Chem. Rev., 2022, 122, 8261–8312 CrossRef CAS PubMed.
- S. J. Geier, C. M. Vogels, J. A. Melanson and S. A. Westcott, The transition metal-catalysed hydroboration reaction, Chem. Soc. Rev., 2022, 51, 8877–8922 RSC.
- C. Jensen, The advantages and difficulties of recent work in molecular catalysis, J. Thermodyn. Catal., 2022, 13, 1–2 Search PubMed.
- S. K. Bose, L. Mao, L. Kuehn, U. Radius, J. Nekvinda, W. L. Santos, S. A. Westcott, P. G. Steel and T. B. Marder, First-Row d-Block Element-Catalyzed Carbon–Boron Bond Formation and Related Processes, Chem. Rev., 2021, 121, 13238–13341 CrossRef CAS PubMed.
- M. Rauch, S. Ruccolo and G. Parkin, Synthesis, Structure, and Reactivity of a Terminal Magnesium Hydride Compound with a Carbatrane Motif, [TismPriBenz]MgH: A Multifunctional Catalyst for Hydrosilylation and Hydroboration, J. Am. Chem. Soc., 2017, 139, 13264–13267 CrossRef CAS PubMed.
- D. Shakhman, A. Dmitrienko, M. Pilkington and G. I. Nikonov, Selective Zinc-Catalyzed 1,2-hydroboration of N-heteroaromatics via a Non–Hydride Mechanism, Eur. J. Inorg. Chem., 2023, 26, e202300293 CrossRef CAS.
- D. Mukherjee, S. Shirase, T. P. Spaniol, K. Mashima and J. Okuda, Magnesium hydridotriphenylborate [Mg(thf)6][HBPh3]2: a versatile hydroboration catalyst, Chem. Commun., 2016, 52, 13155–13158 RSC.
- C. Ginés, B. Parra-Cadenas, R. Fernández-Galán, D. García-Vivó, D. Elorriaga, A. Ramos and F. Carrillo-Hermosilla, Magnesium and Calcium Bisguanidinates: Catalytic Activity in the Hydroboration of Unsaturated Molecules, Adv. Synth. Catal., 2025, 367, e202400843 CrossRef.
- S. Ataie and R. T. Baker, Comparing B–H Bond Activation in NiIIX(NNN)-Catalyzed Nitrile Dihydroboration (X = Anionic N-, C-, O-, S-, or P-donor), Inorg. Chem., 2022, 61, 19998–20007 CrossRef CAS PubMed.
- J. C. Babón, M. A. Esteruelas, I. Fernández, A. M. López and E. Oñate, Dihydroboration of Alkyl Nitriles Catalyzed by an Osmium-Polyhydride: Scope, Kinetics, and Mechanism, Organometallics, 2020, 39, 3864–3872 CrossRef.
- I. Banerjee, S. Anga, K. Bano and T. K. Panda, Efficient and chemoselective hydroboration of organic nitriles promoted by TiIV catalyst supported by unsymmetrical acenaphthenequinonediimine ligand, J. Organomet. Chem., 2019, 902, 120958 CrossRef CAS.
- A. R. Bazkiaei, M. Wiseman and M. Findlater, Iron-catalysed hydroboration of non-activated imines and nitriles: kinetic and mechanistic studies, RSC Adv., 2021, 11, 15284–15289 RSC.
- H. Ben-Daat, C. L. Rock, M. Flores, T. L. Groy, A. C. Bowman and R. J. Trovitch, Hydroboration of alkynes and nitriles using an α-diimine cobalt hydride catalyst, Chem. Commun., 2017, 53, 7333–7336 RSC.
- J. Bhattacharjee, A. Harinath, K. Bano and T. K. Panda, Highly Chemoselective Hydroboration of Alkynes and Nitriles Catalyzed by Group 4 Metal Amidophosphine–Borane Complexes, ACS Omega, 2020, 5, 1595–1606 CrossRef CAS PubMed.
- S. Caddick, D. B. Judd, A. K. d. K. Lewis, M. T. Reich and M. R. V. Williams, A generic approach for the catalytic reduction of nitriles, Tetrahedron, 2003, 59, 5417–5423 CrossRef CAS.
- S. Das, J. Maity and T. K. Panda, Metal/Non-Metal Catalyzed Activation of Organic Nitriles, Chem. Rec., 2022, 22, e202200192 CrossRef CAS PubMed.
- J. B. Geri and N. K. Szymczak, A Proton-Switchable Bifunctional Ruthenium Complex That Catalyzes Nitrile Hydroboration, J. Am. Chem. Soc., 2015, 137, 12808–12814 CrossRef CAS PubMed.
- C. Ghosh, S. Kim, M. R. Mena, J.-H. Kim, R. Pal, C. L. Rock, T. L. Groy, M.-H. Baik and R. J. Trovitch, Efficient Cobalt Catalyst for Ambient-Temperature Nitrile Dihydroboration, the Elucidation of a Chelate-Assisted Borylation Mechanism, and a New Synthetic Route to Amides, J. Am. Chem. Soc., 2019, 141, 15327–15337 CrossRef CAS PubMed.
- K. A. Gudun, A. Slamova, D. Hayrapetyan and A. Y. Khalimon, Efficient Co-Catalyzed Double Hydroboration of Nitriles: Application to One-Pot Conversion of Nitriles to Aldimines, Chem. – Eur. J., 2020, 26, 4963–4968 CrossRef CAS PubMed.
- A. D. Ibrahim, S. W. Entsminger and A. R. Fout, Insights into a Chemoselective Cobalt Catalyst for the Hydroboration of Alkenes and Nitriles, ACS Catal., 2017, 7, 3730–3734 CrossRef CAS.
- M. Itazaki and H. Nakazawa, Selective Double Addition Reaction of an E-H Bond (E = Si, B) to a C
N Triple Bond of Organonitriles, Molecules, 2018, 23, 2769 CrossRef PubMed. - A. Kaithal, B. Chatterjee and C. Gunanathan, Ruthenium-Catalyzed Selective Hydroboration of Nitriles and Imines, J. Org. Chem., 2016, 81, 11153–11161 CrossRef CAS PubMed.
- A. Y. Khalimon, P. Farha, L. G. Kuzmina and G. I. Nikonov, Catalytic hydroboration by an imido-hydrido complex of Mo(iv), Chem. Commun., 2012, 48, 455–457 RSC.
- T. Kitano, T. Komuro and H. Tobita, Double and Single Hydroboration of Nitriles Catalyzed by a Ruthenium–Bis(silyl)xanthene Complex: Application to One-Pot Synthesis of Diarylamines and N-Arylimines, Organometallics, 2019, 38, 1417–1420 CrossRef CAS.
- G. Nakamura, Y. Nakajima, K. Matsumoto, V. Srinivas and S. Shimada, Nitrile hydroboration reactions catalysed by simple nickel salts, bis(acetylacetonato)nickel(II) and its derivatives, Catal. Sci. Technol., 2017, 7, 3196–3199 RSC.
- T. T. Nguyen, J.-H. Kim, S. Kim, C. Oh, M. Flores, T. L. Groy, M.-H. Baik and R. J. Trovitch, Scope and mechanism of nitrile dihydroboration mediated by a β-diketiminate manganese hydride catalyst, Chem. Commun., 2020, 56, 3959–3962 RSC.
- V. K. Pandey, C. S. Tiwari and A. Rit, Silver-Catalyzed Hydroboration of C–X (X = C, O, N) Multiple Bonds, Org. Lett., 2021, 23, 1681–1686 CrossRef CAS PubMed.
- Z. Huang, S. Wang, X. Zhu, Q. Yuan, Y. Wei, S. Zhou and X. Mu, Well-Defined Amidate-Functionalized N-Heterocyclic Carbene -Supported Rare-Earth Metal Complexes as Catalysts for Efficient Hydroboration of Unactivated Imines and Nitriles, Inorg. Chem., 2018, 57, 15069–15078 CrossRef CAS PubMed.
- S. Pradhan, R. V. Sankar and C. Gunanathan, A Boron-Nitrogen Double Transborylation Strategy for Borane-Catalyzed Hydroboration of Nitriles, J. Org. Chem., 2022, 87, 12386–12396 CrossRef CAS PubMed.
- B. Yan, X. He, C. Ni, Z. Yang and X. Ma, n-Butyllithium Catalyzed Hydroboration of Nitriles and Carbodiimides, ChemCatChem, 2021, 13, 851–854 CrossRef CAS.
- P. Ghosh and A. J. von Wangelin, Lithium amide catalyzed hydroboration of nitriles, Org. Chem. Front., 2020, 7, 960–966 RSC.
- D. Bedi, A. Brar and M. Findlater, Transition metal- and solvent-free double hydroboration of nitriles, Green Chem., 2020, 22, 1125–1128 RSC.
- W. Liu, Y. Ding, D. Jin, Q. Shen, B. Yan, X. Ma and Z. Yang, Organic aluminum hydrides catalyze nitrile hydroboration, Green Chem., 2019, 21, 3812–3815 RSC.
- A. Harinath, J. Bhattacharjee and T. K. Panda, Catalytic Hydroboration of Organic Nitriles Promoted by Aluminum Complex, Adv. Synth. Catal., 2019, 361, 850–857 CrossRef CAS.
- Y. Ding, X. Ma, Y. Liu, W. Liu, Z. Yang and H. W. Roesky, Alkylaluminum Complexes as Precatalysts in Hydroboration of Nitriles and Carbodiimides, Organometallics, 2019, 38, 3092–3097 CrossRef CAS.
- J. Li, M. Luo, X. Sheng, H. Hua, W. Yao, S. A. Pullarkat, L. Xu and M. Ma, Unsymmetrical β-diketiminate magnesium(I) complexes: syntheses and application in catalytic hydroboration of alkyne, nitrile and carbonyl compounds, Org. Chem. Front., 2018, 5, 3538–3547 RSC.
- C. Weetman, M. D. Anker, M. Arrowsmith, G. Kociok-Köhn, D. J. Liptrot, M. F. Mahon and M. S. Hill, Magnesium-catalysed nitrile hydroboration, Chem. Sci., 2016, 7, 628–641 RSC.
- M. Arrowsmith, M. S. Hill and G. Kociok-Köhn, Magnesium Catalysis of Imine Hydroboration, Chem. – Eur. J., 2013, 19, 2776–2783 CrossRef CAS PubMed.
- D. Haddenham, L. Pasumansky, J. DeSoto, S. Eagon and B. Singaram, Reductions of Aliphatic and Aromatic Nitriles to Primary Amines with Diisopropylaminoborane, J. Org. Chem., 2009, 74, 1964–1970 CrossRef CAS PubMed.
- X. Wang and X. Xu, Hydroboration of nitriles and imines by highly active zinc dihydride catalysts, RSC Adv., 2021, 11, 1128–1133 RSC.
- S. Das, J. Bhattacharjee and T. K. Panda, An imidazolin-2-iminato ligand organozinc complex as a catalyst for hydroboration of organic nitriles, New J. Chem., 2019, 43, 16812–16818 RSC.
- N. Sarkar, S. Bera and S. Nembenna, Aluminum-Catalyzed Selective Hydroboration of Nitriles and Alkynes: A Multifunctional Catalyst, J. Org. Chem., 2020, 85, 4999–5009 CrossRef CAS PubMed.
- C. Weetman, M. S. Hill and M. F. Mahon, Magnesium-catalysed hydroboration of isonitriles, Chem. Commun., 2015, 51, 14477–14480 RSC.
- R. K. Sahoo, S. Rajput, S. Dutta, K. Sahu and S. Nembenna, Zinc Hydride-Catalyzed Dihydroboration of Isonitriles and Nitriles: Mechanistic Studies with the Structurally Characterized Zinc Intermediates, Organometallics, 2023, 42, 2293–2303 CrossRef CAS.
- S. Ataie, J. S. Ovens and R. T. Baker, Solvent-free Zn (NSNO) complex-catalysed dihydroboration of nitriles, Chem. Commun., 2022, 58, 8266–8269 RSC.
- S. Saha and M. S. Eisen, Catalytic Recycling of a Th–H Bond via Single or Double Hydroboration of Inactivated Imines or Nitriles, ACS Catal., 2019, 9, 5947–5956 CrossRef CAS.
- Y. Sarazin and P. M. Chapple, in Comprehensive Organometallic Chemistry IV, ed. G. Parkin, K. Meyer and D. O'hare, Elsevier, Oxford, 2022, pp. 104–192 Search PubMed.
- S. Nembenna, N. Sarkar, R. K. Sahoo and S. Mukhopadhyay, in Comprehensive Organometallic Chemistry IV, ed. G. Parkin, K. Meyer and D. O'hare, Elsevier, Oxford, 2022, pp. 71–241 Search PubMed.
- M. S. Hill, D. J. Liptrot and C. Weetman, Alkaline earths as main group reagents in molecular catalysis, Chem. Soc. Rev., 2016, 45, 972–988 RSC.
- M. R. Crimmin and M. S. Hill, in Alkaline-Earth Metal Compounds: Oddities and Applications, ed. S. Harder, Springer Berlin Heidelberg, Berlin, Heidelberg, 2013, pp. 191–241 Search PubMed.
- S. Harder, From Limestone to Catalysis: Application of Calcium Compounds as Homogeneous Catalysts, Chem. Rev., 2010, 110, 3852–3876 CrossRef CAS PubMed.
- A. Stasch, in Catalysis with Earth-abundant Elements, ed. U. Schneider and S. Thomas, The Royal Society of Chemistry, 2020, pp. 55–80. 10.1039/9781788012775-00055.
- M. J. Evans and C. Jones, Low oxidation state and hydrido group 2 complexes: synthesis and applications in the activation of gaseous substrates, Chem. Soc. Rev., 2024, 53, 5054–5082 RSC.
- S. Mukhopadhyay, S. Rajput, R. K. Sahoo and S. Nembenna, Insights Into the Magnesium-Catalyzed C–C Coupling of Terminal Alkynes with Carbodiimides, Eur. J. Org. Chem., 2024, e202400702 CrossRef CAS.
- T. Peddarao, A. Baishya, N. Sarkar, R. Acharya and S. Nembenna, Conjugated Bis-Guanidines (CBGs) as β-Diketimine Analogues: Synthesis, Characterization of CBGs/Their Lithium Salts and CBG Li Catalyzed Addition of B–H and TMSCN to Carbonyls, Eur. J. Inorg. Chem., 2021, 2021, 2034–2046 CrossRef CAS.
- R. K. Sahoo, N. Sarkar and S. Nembenna, Zinc Hydride Catalyzed Chemoselective Hydroboration of Isocyanates: Amide Bond Formation and C=O Bond Cleavage, Angew. Chem., Int. Ed., 2021, 60, 11991–12000 CrossRef CAS PubMed.
- M. Westerhausen, Synthesis and spectroscopic properties of bis(trimethylsilyl)amides of the alkaline-earth metals magnesium, calcium, strontium, and barium, Inorg. Chem., 1991, 30, 96–101 CrossRef CAS.
- C. Bakewell, Magnesium hydrides bearing sterically demanding amidinate ligands: synthesis, reactivity and catalytic application, Dalton Trans., 2020, 49, 11354–11360 RSC.
- J. Spielmann and S. Harder, Hydrocarbon-Soluble Calcium Hydride: A “Worker-Bee” in Calcium Chemistry, Chem. – Eur. J., 2007, 13, 8928–8938 CrossRef CAS PubMed.
- J. Brettar and S. Harder, Rational Design of a Well-Defined Soluble Calcium Hydride Complex, Angew. Chem., Int. Ed., 2006, 45, 3474–3478 CrossRef PubMed.
- A. Causero, G. Ballmann, J. Pahl, C. Färber, J. Intemann and S. Harder, β-Diketiminate calcium hydride complexes: the importance of solvent effects, Dalton Trans., 2017, 46, 1822–1831 RSC.
- P. Rinke, H. Görls, P. Liebing and R. Kretschmer, Synthesis and Characterization of Magnesium-Hydride Complexes with Alkylene-Bridged Bis(β-diketiminate) Ligands, Organometallics, 2023, 42, 2287–2292 CrossRef CAS.
- P. Rinke, H. Görls and R. Kretschmer, Calcium and Magnesium Bis(β-diketiminate) Complexes: Impact of the Alkylene Bridge on Schlenk-Type Rearrangements, Inorg. Chem., 2021, 60, 5310–5321 CrossRef CAS PubMed.
- M. P. Coles, The role of the bis-trimethylsilylamido ligand, [N{SiMe3}2]−, in main group chemistry. Part 1: Structural chemistry of the s-block elements, Coord. Chem. Rev., 2015, 297–298, 2–23 CrossRef CAS.
- A. D. Bage, K. Nicholson, T. A. Hunt, T. Langer and S. P. Thomas, The Hidden Role of Boranes and Borohydrides in Hydroboration Catalysis, ACS Catal., 2020, 10, 13479–13486 CrossRef CAS.
- A. D. Bage, T. A. Hunt and S. P. Thomas, Hidden Boron Catalysis: Nucleophile-Promoted Decomposition of HBpin, Org. Lett., 2020, 22, 4107–4112 CrossRef CAS PubMed.
- S. Mukhopadhyay, A. G. Patro, R. S. Vadavi and S. Nembenna, Coordination Chemistry of Main Group Metals with Organic Isocyanides, Eur. J. Inorg. Chem., 2022, 2022, e202200469 CrossRef CAS.
- A. P. Dove, P. Hormnirun, E. L. Marshall, J. A. Segal, A. J. P. White, D. J. Williams and V. C. Gibson, Low coordinate magnesium chemistry supported by a bulky β-diketiminate ligand, Dalton Trans., 2003, 3088–3097, 10.1039/B303550F.
- S. Itsuno, C. Hachisuka, K. Kitano and K. Ito, Synthesis of secondary carbinamine via n-boryl imines generated from nitriles and alkylboranes, Tetrahedron Lett., 1992, 33, 627–630 CrossRef CAS.
- H. C. Brown, P. Heim and N. M. Yoon, Selective reductions. XV. Reaction of diborane in tetrahydrofuran with selected organic compounds containing representative functional groups, J. Am. Chem. Soc., 1970, 92, 1637–1646 CrossRef CAS.
- F. Meger, A. C. W. Kwok, F. Gilch, D. R. Willcox, A. J. Hendy, K. Nicholson, A. D. Bage, T. Langer, T. A. Hunt and S. P. Thomas, B–N/B–H Transborylation: borane-catalysed nitrile hydroboration, Beilstein J. Org. Chem., 2022, 18, 1332–1337 CrossRef CAS PubMed.
- N. Matsumi and Y. Chujo, Synthesis of π-Conjugated Poly(cyclodiborazane)s by Organometallic Polycondensation, Macromolecules, 2000, 33, 8146–8148 CrossRef CAS.
- Y. Chujo, I. Tomita and T. Saegusa, Hydroboration Polymerization of Dicyano Compounds. 4. Synthesis of Stable Poly(cyclodiborazane)s from Dialkylboranes, Macromolecules, 1994, 27, 6714–6717 CrossRef CAS.
- M. Yalpani, R. Köster and R. Boese, Monomeric Aldimino-diorganoboranes and Aspects of Their Chemistry, Chem. Ber., 1993, 126, 285–288 CrossRef CAS.
- U. E. Diner, M. Worsley, J. W. Lown and J.-A. Forsythe, A novel asymmetric synthesis of α-aminoacids from nitriles employing diisopinocampheylborane as a chiral agent, Tetrahedron Lett., 1972, 13, 3145–3148 CrossRef.
- M. F. Hawthorne, Amine boranes—X: Alkylideneamino t-butylboranes. The hydroboration of nitriles with trimethylamine t-butylborane, Tetrahedron, 1962, 17, 117–122 CrossRef CAS.
- B. Okokhere-Edeghoghon, S. E. Neale, M. S. Hill, M. F. Mahon and C. L. McMullin, Isocyanate deoxygenation by a molecular magnesium silanide, Dalton Trans., 2022, 51, 136–144 RSC.
- J. Burés, Variable Time Normalization Analysis: General Graphical Elucidation of Reaction Orders from Concentration Profiles, Angew. Chem., Int. Ed., 2016, 55, 16084–16087 CrossRef PubMed.
- M. R. Crimmin, M. Arrowsmith, I. J. Casely, P. A. Procopiou, A. G. M. Barrett and M. S. Hill, Intramolecular Hydroamination of Aminoalkenes by Calcium and Magnesium Complexes: A Synthetic and Mechanistic Study, J. Am. Chem. Soc., 2009, 131, 9670–9685 CrossRef CAS PubMed.
- M. R. Crimmin, I. J. Casely and M. S. Hill, Calcium-Mediated Intramolecular Hydroamination Catalysis, J. Am. Chem. Soc., 2005, 127, 2042–2043 CrossRef CAS PubMed.
- M. Arrowsmith, M. R. Crimmin, A. G. M. Barrett, G. Kociok-Köhn, P. A. Procopiou and M. S. Hill, Cation Charge Density and Precatalyst Selection in Group 2-Catalyzed Aminoalkene Hydroamination, Organometallics, 2011, 30, 1493–1506 CrossRef CAS.
- K. A. Johnson and R. S. Goody, The Original Michaelis Constant: Translation of the 1913 Michaelis–Menten Paper, Biochemistry, 2011, 50, 8264–8269 CrossRef CAS PubMed.
- L. Michaelis and M. L. Menten, Die kinetik der invertinwirkung, Biochem. Z., 1913, 49, 352 Search PubMed.
- Y. Xia, H. Jiang and W. Wu, Recent Advances in Chemical Modifications of Nitriles, Eur. J. Inorg. Chem., 2021, 2021, 6658–6669 CrossRef CAS.
-
(a) CCDC 2476567: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2p4286;
(b) CCDC 2476571: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2p42db.
|
| This journal is © the Partner Organisations 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.