Sterically hindered P,N-type amidophosphines {{(o-PPh2)C6H4}C(O)NH(R)}: synthesis, transition metal chemistry, and catalytic activity in stereoselective Heck coupling of aryl chlorides

Gazal Sabharwal , Khilesh C. Dwivedi and Maravanji S. Balakrishna *
Phosphorus Laboratory, Department of Chemistry, Indian Institute of Technology Bombay, Powai, Mumbai 400076, India. E-mail: krishna@chem.iitb.ac.in; msb_krishna@iitb.ac.in; Fax: +91-22-5172-3480/2576-7152

Received 9th June 2025 , Accepted 9th July 2025

First published on 11th July 2025


Abstract

This paper reports the synthesis, coordination chemistry, and catalytic applications of two amide-based monophosphine ligands, {{(o-PPh2)C6H4}C(O)N(H)(C12H18)} (L1) and {{(o-PPh2)C6H4}C(O)NH(C33H28)} (L2). Ligand L1, upon reaction with [Pd(COD)Cl2] and [Pd(OAc)2], yielded dimeric P,N-coordinated palladium complexes [{(PdCl)2}{(o-PPh2)C6H4C(O)N(C12H18)}22-P,N] (1) and [{(Pd(OAc))2}{(o-PPh2)C6H4C(O)N(C12H18)}22-P,N] (2), respectively. Further treatment of L1 with CuI and AgBr afforded monodentate η1-P coordinated copper and silver complexes, [{(CuI)2}{(o-PPh2)C6H4C(O)NH(C12H18)}21-P] (3) and [{(AgBr)2}{(o-PPh2)C6H4C(O)N(H)(C12H18)}21-P] (4). Reactions of L1 with AgBF4 and AgClO4 led to the formation of tridentate P,O,C-coordinated silver complexes, namely [{(AgX)2}{(o-PPh2)C6H4C(O)N(H)(C12H18)}23-P,O,C], where X = BF4 (5) and ClO4 (6). Ligand L2 reacted with CuI to form a monodentate P-coordinated CuI complex, [{(CuI)2}{(o-PPh2)C6H4C(O)NH(C33H28)}21-P] (8), and with [Ru(p-cymene)Cl2]2 to afford a P,O-chelated RuII complex, [{Ru(p-cymene)Cl}{(o-PPh2)C6H4C(O)NH(C33H28)}-κ2-P,O]PF6 (9). Notably, the in situ-generated Pd nanoparticles derived from complex 1, stabilized by the P,N ligand, exhibited excellent catalytic performance in Heck cross-coupling reactions of aryl chlorides with styrene derivatives. These reactions proceed under mild conditions, providing trans-stilbene products in high yields (90–99%) with low catalyst loading and good functional group tolerance.


Introduction

Ligand design is central to advancing coordination chemistry and transition metal catalysis, with tertiary phosphines (PR3) standing out for their ability to finely tune the electronic and steric environment of metal centres.1,2 Phosphine ligands incorporating nitrogen donor groups (P,N-ligands) are particularly valuable due to their hemilabile nature, combining soft phosphorus and hard nitrogen donors.3–7 This duality allows for dynamic coordination behaviour: the nitrogen donor can reversibly dissociate to create vacant coordination sites, while the chelating interaction stabilizes the complex in the absence of substrate.8–11 The performance of P,N-ligands is further enhanced by their electronic asymmetry and steric versatility.12 Incorporating additional heteroatoms or sterically demanding substituents increases both the stability and reactivity of metal complexes, enabling efficient catalysis across a broad range of transformations.13–19 Notably, phosphines bearing pendant aromatic groups adjacent to phosphorus exhibit diverse coordination modes (η2–η6), contingent on the metal's electronic state, thus offering further tunability in catalytic systems.20–24 In particular, amide-derived phosphines present an attractive platform, owing to the inherent coordinating properties of the amide moiety and the ease of structural modification at the nitrogen center, which facilitates the incorporation of diverse steric and electronic features.25 Despite extensive studies on conventional P,N- and P,O-type ligands,26–28 the class of aromatic amide-based monophosphines featuring dangling tertiary amide groups remains relatively underexplored. The introduction of dynamic steric hindrance via bulky N-aryl or N-alkyl substituents provides an opportunity to modulate ligand flexibility and metal coordination geometry, potentially facilitating key steps in catalytic cycles such as substrate coordination, oxidative addition, and reductive elimination.29 While related P,N-ligands and amido-phosphines have been reported in the literature,30–32 the use of aromatic amide-derived monophosphines with bulky N-substituents and their ability to support a range of coordination modes (P,P,N, P,O,C) across multiple metals (Pd, Cu, Ag, Ru) remains underexplored.

Palladium plays a central role in transition metal catalysis due to its broad functional group tolerance, mild reaction conditions, and capacity to activate a variety of bonds, including C–H, C–X (X = Cl, Br, I), and C–Sn.33–37 Its catalytic efficiency is markedly improved through coordination with electron-rich ligands, especially phosphines, which enhance the electron density at the metal center and promote key steps in the catalytic cycle.38–40

Palladium nanoparticles (Pd NPs) stabilized by various ligands under mild conditions have garnered significant attention in recent years due to their catalytic potential.41–45 Phosphorus-based ligands, in particular, have been employed as surfactants to effectively stabilize Pd NPs.17,46–48 Notably, in situ-generated palladium nanoparticles prepared under milder and less hazardous conditions have demonstrated superior catalytic activity compared to their corresponding molecular complexes, while also offering a more cost-effective alternative.43,49–52

The palladium-catalyzed arylation of olefins with aryl halides, widely known as the Heck reaction, has garnered considerable interest over the past decade for its broad applicability in carbon–carbon bond formation.53–61 It is pivotal in synthesizing key intermediates for the pharmaceutical and chemical industries.62,63 Beyond conventional aryl halides, alternative aryl sources such as aryl triflates, diazonium salts, sulfonyl halides, and aroyl halides, have been explored, along with the development of highly efficient catalytic systems.64–67 However, industrial application remains limited, particularly with chloroarenes, despite their low cost and availability. The Heck coupling reaction continues to face significant challenges, including high catalyst loadings (1–5 mol%), elevated reaction temperatures (>130 °C), dependence on additives, reducing agents, or phase-transfer catalysts, as well as limitations in selectivity and overall efficiency, underscoring the need for more active, selective, and practical catalytic systems.68,69

Sasson and co-workers, reported significant side reactions, including double bond hydrogenation, styrene reduction to ethylbenzene, homocoupling, and hydrodehalogenation, when using equimolar haloarene and styrene with high loading of Pd/C, PEG-400, sodium formate, and base.70 Xu and co-workers achieved trans-Heck products from aryl chlorides and olefins using n-Bu4N+OAc as base, but only under high palladium and ligand loadings.71 Beller and co-workers developed a cyclometallated PdII complex that enabled coupling of aryl bromides with n-butyl acrylate at low catalyst loadings; however, the method required elevated temperatures and prolonged reaction times (Scheme 1).72 Recently Madrahimov and co-workers, have demonstrated the heterogenization of mono(phosphine)-Pd complexes on UiO-66 MOF surfaces for Heck reactions, achieving moderate yields (40–92%) but requiring high catalyst loadings.73 Similarly, Gevorgyan and co-workers developed a visible-light-induced Pd-catalyzed Heck coupling of oximes with alkyl halides, though the method suffered from high catalyst loadings, limited scope (ineffective for chloro derivatives), and formation of cis/trans mixtures.74 Strassner and co-workers employed tunable aryl alkyl ionic liquids (TAAILs) at elevated temperatures (140 °C), with high Pd loadings, to couple bromobenzene and styrene, again producing trans-stilbene.75


image file: d5dt01353d-s1.tif
Scheme 1 Heck-coupling reaction reported by various groups.

To overcome existing limitations in Heck coupling, we present a novel amido-phosphine ligand and its dimeric P,N-coordinated PdII complex, which efficiently catalyses the arylation of styrene derivatives with aryl chlorides under mild conditions. Operating at low catalyst loadings, the system delivers trans-stilbene products in excellent yields (90–99%) with broad substrate scope and high selectivity. This study demonstrates how the hemilabile, flexible nature of the ligand framework can be exploited to design efficient and sustainable catalysts for C–C bond-forming transformations.

Results and discussion

Synthesis of amide based monophosphine ligand: N-(2,6-diisopropylphenyl)-2-(diphenylphosphaneyl)benzamide (L1)

The reaction of 2,6-diisopropylaniline with 2-bromobenzoyl chloride in 1[thin space (1/6-em)]:[thin space (1/6-em)]1 molar ratio afforded the corresponding amide in quantitative yield. Subsequent treatment with 1.3 equivalents of diphenylphosphine in the presence of catalytic amount of [Pd(PPh3)4] and K2CO3 furnished the monophosphine ligand L1 in high yield (Scheme 2).76,77 The structure of L1 was confirmed by NMR, mass, IR spectroscopy, and single-crystal X-ray analysis. In the 1H NMR spectrum, the NH proton appeared at 7.95 ppm, while the 13C{1H} NMR spectrum showed the amide carbonyl carbon at 168.8 ppm. Aromatic proton resonances were observed in the range of 7.00–7.80 ppm. The IR spectrum exhibited characteristic absorptions at 3364 cm−1 (N–H stretch) and 1653 cm−1 (C[double bond, length as m-dash]O stretch). Crystals suitable for X-ray diffraction were obtained from dichloromethane, and the molecular structure is depicted in Fig. 1, with selected bond lengths (Å) and angles (°) mentioned in the figure caption.
image file: d5dt01353d-s2.tif
Scheme 2 Synthesis of L1.

image file: d5dt01353d-f1.tif
Fig. 1 Single X-ray crystal structure of ligand L1. All hydrogen atoms have been omitted for clarity. Displacement ellipsoids are drawn at 50% probability level. Selected bond lengths [Å] and bond angles [°]: P1–C1 1.868(2), C7–O1 1.236(2), N1–C7 1.355(3), NI–C7–O1 121.8(2).

Synthesis of PdII, CuI and AgI complexes

The reactions of ligand L1 with [Pd(COD)Cl2] and [Pd(OAc)2] in 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio, in dichloromethane afforded the complexes [{(PdCl)2}{{(o-PPh2)C6H4}C(O)N(C12H18)}22-P,N] (1) and [{(CH3C(O)O)(Pd)2}{{(o-PPh2)C6H4}C(O)N(C12H18)}22-P,N] (2) (Scheme 3) in good yields as intense yellow colour solids. In the IR spectra of both the complexes the N–H stretching frequency disappears. In case of complex 2, carbonyl stretching frequency appeared at 1549 cm−1 and the C–O stretching frequency appears at 1099 cm−1. The crystal structures of both the complexes are shown in Fig. 2, with selected bond lengths and bond angles mentioned in Table 1. Complexes 1 and 2 are both dinuclear, with palladium centers adopting distorted square planar geometries. Complex 1 features chloride bridges, forming two six-membered and one four-membered metallacycles, whereas complex 2 contains acetate bridges, leading to the formation of two six-membered and one eight-membered metallacycles.
image file: d5dt01353d-s3.tif
Scheme 3 Synthesis of PN-PdII complexes 1 and 2.

image file: d5dt01353d-f2.tif
Fig. 2 Single X-ray crystal structure of complexes 1 and 2. All hydrogen atoms have been omitted for clarity. Displacement ellipsoids are drawn at 50% probability level.
Table 1 Selected bond lengths (Å) and bond angles (°) of 1–2
  1 2
P1–Pd1 2.2096(14) 2.1787(8)
N1–Pd1 2.040(4) 2.014(2)
Pd1–Cl1 2.4549(14)
Pd1–Cl1′ 2.3491(13)
Pd1–O3 2.0413(18)
Pd1–O6 2.158(2)
N1–C7 1.337(6) 1.348(4)
C7–O1 1.231(6) 1.243(3)
O5–C65 1.280(3)
O3–C63 1.271(3)
Pd1–P1–N1 89.81(12) 88.87(7)
Cl1–Pd1–Cl1′ 84.57(5)
N1–Pd1–Cl1 93.23(12)
P1–Pd1–Cl1 176.25(5)
O3–Pd1–P1 90.91(6)
N1–Pd1–O3 175.43(8)
O6–Pd1–P1 176.35(6)
N1–Pd1–O6 94.54(8)


The reaction of ligand L1 with CuI in 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio in dichloromethane/acetonitrile mixture and with AgBr in dichloromethane afforded the corresponding mononuclear phosphine complexes [{(CuI)2}{{(o-PPh2)C6H4}C(O)NH(C12H18)}21-P] (3) and [{(AgBr)2}{{(o-PPh2)C6H4}C(O)N(H)(C12H18)}21-P] (4), in good yields. Further treatment of L1 with AgBF4 and AgClO4 in dichloromethane, in 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio, yielded the complexes [{(AgX)2}{{(o-PPh2)C6H4}C(O)N(H)(C12H18)}23-P,O,C] (X = BF4 (5), ClO4 (6)) as white solids in good yields (Scheme 4). The IR spectra showed N–H stretching frequencies at 3284 cm−1 (5) and 3248 cm−1 (6) and a common carbonyl stretching frequency at 1609 cm−1. The structures of complexes 3–6 were unambiguously confirmed by single-crystal X-ray diffraction.


image file: d5dt01353d-s4.tif
Scheme 4 Synthesis of CuI and AgI dimeric complexes 3–6.

Perspective views are shown in Fig. 3, selected bond lengths and bond angles are mentioned in Table 2 for better comparison. Complexes 3–6 contain CuI and AgI centers, each adopting a distorted trigonal planar geometry. Complexes 3 and 4 are neutral species, while 5 and 6 are cationic, balanced by ClO4 and BF4 counterions, respectively. All four complexes exhibit tridentate coordination. In 3 and 4, the ligands coordinate through an η1-P donor mode, accompanied by bridging halide ions, leading to the formation of four-membered metallacycles. In contrast, complexes 5 and 6 display P,O,C-tridentate coordination, resulting in the formation of two five-membered and one twelve-membered metallacycles.


image file: d5dt01353d-f3.tif
Fig. 3 Single X-ray crystal structure of complexes 3–6. All hydrogen atoms have been omitted for clarity. Displacement ellipsoids are drawn at 50% probability level.
Table 2 Selected bond lengths (Å) and bond angles (°) of 3–6
  3 (X = I) (M = Cu) 4 (X = Br) (M = Ag) 5 (M = Ag) 6 (M = Ag)
P1–M1 2.2257(6) 2.4154(7) 2.3669(11) 2.3688(6)
M1–X1 2.5535(3) 2.6947(6)
M1′–X1 2.5847(3) 2.6758(5)
C8–M1 2.629(4) 2.633(2)
O1–M1 2.271(3) 2.2707(19)
M1–M1′ 2.7235(5) 3.6989(10)
N1–C7 1.344(3) 1.357(3) 1.337(6) 1.332(3)
O1–C7 1.227(3) 1.230(3) 1.246(6) 1.247(3)
M1–X1–M1′ 64.015(11) 87.062(16)
X1–M1–X1′ 115.987(11) 92.938(16)
P1–M1–X1 127.309(18) 134.247 (18)
P1–M1–X1′ 115.794(18) 134.606(17)
P1–M1–C8 148.05(10) 147.92(5)
O1–M1–P1 142.96(9) 142.99(5)


Synthesis of amide based monophosphine ligand: N-(2,6-dibenzhydryl-4-methylphenyl)-2-(diphenylphosphaneyl)benzamide (L2)

The dialkylation of p-toluidine using diphenylmethanol as alkylating agent under solvent-free conditions, in stoichiometric amount of conc. HCl and ZnCl2 produced 2,6-dibenzhydryl-4-methylaniline, which upon further treatment with 2-bromobenzoyl chloride in 1[thin space (1/6-em)]:[thin space (1/6-em)]1 molar ratio afforded new amide derivative A. The treatment of amide derivative with 1.3 equiv. of diphenylphosphine in presence of catalytic amount of [Pd(PPh3)4] and K2CO3 yielded new monophosphine ligand L2 in high yield. In 1H NMR spectrum, characteristics NH proton appeared at 7.29 ppm and in 13C{1H} NMR spectrum carbonyl carbon appeared at 168.2 ppm. Aromatic protons showed resonance around 7–7.8 ppm. In IR spectrum the secondary amide band appeared at 3401 cm−1 and the carbonyl stretching frequency appeared at 1684 cm−1. Treatment of L2 with H2O2 in THF resulted in phosphine oxide 7 (Scheme 5). The structure 7 was confirmed by single crystal X-ray analysis and is shown in Fig. 4, with selected bond lengths and bond angles mentioned in the figure captions.
image file: d5dt01353d-s5.tif
Scheme 5 Synthesis of ligand L2 and 7.

image file: d5dt01353d-f4.tif
Fig. 4 Molecular structure of 7. All hydrogen atoms and solvent molecules have been omitted for clarity. Displacement ellipsoids are drawn at 50% probability level. Bond length(Å) and bond angles (°): P1–O2 1.4738(17), P1–C7 1.823(2), O1–C1 1.220(3), N1–C20 1.435(3), N1–C1 1.350(3), O2–P1–C7 112.28(10), C1–N1–C20 123.57(18), O1–C1–N1 123.5(2), O1–C1–C2 122.12(19), N1–C1–C2 114.25(18).

Synthesis of CuI and RuII complexes

Reaction of ligand L2 with CuI in 1[thin space (1/6-em)]:[thin space (1/6-em)]1 molar ratio in a dichloromethane/acetonitrile mixture yielded complex [{(CuI)2}{{(o-PPh2)C6H4}C(O)NH(C33H28)}21-P] (8) as a yellow solid in good yield. The IR spectrum of 8 exhibited a secondary amide N–H stretching band at 3435 cm−1 and a carbonyl stretching frequency at 1652 cm−1. Treatment of ligand L2 with [Ru(p-cymene)Cl2]2 in the presence of NH4PF6 (1[thin space (1/6-em)]:[thin space (1/6-em)]1 molar ratio) under reflux in methanol afforded the cationic Ru complex [{(Ru(p-cymene)Cl)}{{(o-PPh2)C6H4}C(O)NH(C33H28)}-κ2-P,O](PF6) (9), isolated as an orange solid in good yield (Scheme 6).
image file: d5dt01353d-s6.tif
Scheme 6 Synthesis of complexes 8 and 9.

Both complexes were crystallized by slow diffusion of petroleum ether into dichloromethane solutions of the respective compounds. The molecular structures of complexes 8 and 9 are shown in Fig. 5, with selected bond lengths and angles provided in the figure captions. In complex 8, both CuI centres adopt a distorted trigonal planar geometry. Complex 9 being a catanionic complex features RuII center in a distorted tetrahedral geometry, with PF6 as a counter ion.


image file: d5dt01353d-f5.tif
Fig. 5 Single X-ray crystal structure of complexes 8 and 9. All hydrogen atoms have been omitted for clarity. Displacement ellipsoids are drawn at 50% probability level. Selected bond distances [Å] and bond angles [°]: for 8: Cu1–I1 2.5827(5), Cu1′–I1 2.5588(6), Cu1–Cu1′ 2.6108(8), P1–Cu1 2.2260(9), N1–C7 1.356(4), O1–C7 1.222(3), Cu1–I1–Cu1′ 61.031(15), I1–Cu1–I1′ 118.969(15), I1–Cu1–Cu1′ 59.032(18), I1′–Cu1–Cu1′ 59.937(18), P1–Cu1–I1 123.47(3), P1–Cu1–I1′ 116.90(3), P1–Cu1–Cu1′ 171.16(3). For 9: Ru1–Cl1 2.3868(12), Ru1–P1 2.3308(14), Ru1–O1 2.105(3), O1–C1 1.258(5), P1–Ru1–Cl1 88.05(5), O1–Ru1–Cl1 86.24(9), O1–Ru1–P1 79.69(9).

Structural insights from 31P{1H} NMR spectroscopy

The coordination behaviour of the phosphine moiety in ligands L1 and L2 and their corresponding metal complexes was systematically examined using 31P{1H} NMR spectroscopy. The free ligands L1 and L2 exhibited singlets at −11.4 ppm and −10.6 ppm, respectively, which are characteristic of uncoordinated tertiary phosphines. Upon coordination to PdII, the phosphorus signals shifted markedly downfield, with complex 1 displaying a singlet at 16.6 ppm and complex 2 at 26.3 ppm. These substantial downfield shifts reflect deshielding of the phosphorus nuclei due to coordination with the electron-deficient PdII center, and are consistent with P,N-chelation in a square planar environment.76,77 The further downfield shift observed in complex 2 compared to 1 may arise from the stronger electron-withdrawing effect of the acetate ligands relative to chlorides, enhancing the deshielding of the phosphorus atom.

In contrast, the CuI and AgI complexes derived from L1 (complexes 3–6) exhibited significantly upfield-shifted 31P signals. Complexes 3 and 4 displayed singlets at −6.1 ppm and −8.8 ppm, respectively, consistent with monodentate η1-P coordination of the soft phosphine donor to the d10 CuI and AgI centers.78 The upfield chemical shifts reflect reduced deshielding in the absence of strong π-backbonding and in a less electron-deficient coordination environment. Interestingly, complexes 5 and 6, which involve tridentate P,O,C coordination to AgI, exhibited doublets at 10.9 ppm with 1JP–Ag value of 756.5 and 753.3 Hz, respectively. These doublets confirm direct coordination of phosphorus to the Ag center and reflect a stronger interaction in the more rigid tridentate coordination mode, as well as increased covalency in the P–Ag bond.

The RuII complex 9 derived from L2 exhibited a 31P{1H} singlet at 32.5 ppm, indicating significant deshielding upon P,O-chelation to the electrophilic RuII center in a distorted tetrahedral geometry. Additionally, a septet observed at −144.3 ppm corresponds to the PF6 counterion, with a 1JPF coupling of 712.8 Hz, further confirming the presence of the hexafluorophosphate anion in the complex.

Collectively, these data illustrate that the 31P chemical shifts are highly sensitive to the oxidation state, electron density, and coordination geometry of the metal center, as well as the denticity and rigidity of the ligand framework. Downfield shifts typically accompany coordination to more electron-deficient or π-accepting metals (PdII, RuII), while upfield shifts are observed with softer, less electronegative metal centers (CuI, AgI) in monodentate modes. The presence of P–Ag scalar coupling in complexes 5 and 6 further underscores the strong, covalent nature of the metal–phosphorus interaction in these systems. These results collectively highlight the versatility of amide-functionalized phosphines in modulating electronic environments through varied coordination modes across different metal centers.

Heck-coupling reaction between various aryl chlorides and styrene derivatives promoted by P,N-PdII complex 1

Chlorobenzene and styrene were selected as model substrates to optimize the reaction conditions. No product formation was observed in the absence of either the catalyst or the base (Table 3, entries 1 and 2). Using 0.05 mol% of catalyst 1 afforded an 83% yield of trans-stilbene, which increased to 99% upon increasing the catalyst loading to 0.1 mol% (entries 3 and 4). Among the bases tested, Cs2CO3 proved most effective (entries 4–7), and DMF was identified as the optimal solvent (entries 4 and 8–10). Lowering the reaction temperature to 80 °C resulted in diminished yield (entry 11). The optimal conditions-0.1 mol% catalyst 1, Cs2CO3 as base, and DMF as solvent at 120 °C (entry 4), were adopted for further substrate scope evaluation (Scheme 7).
image file: d5dt01353d-s7.tif
Scheme 7 Substrate scope Heck-coupling reaction. Conditions: aryl chloride (0.50 mmol), styrene (0.60 mmol), Cs2CO3 (0.60 mmol), DMF 2 mL, 120 °C, and catalyst 1 (0.1 mol%). All are isolated yields.
Table 3 Optimization of the reaction condition for Heck-coupling reaction

image file: d5dt01353d-u1.tif

Entry Catalyst Solvent Base Yielda (%)
a Yield determined by GC-MS. NC = no conversion. Chlorobenzene (0.50 mmol), styrene (0.6 mmol), Cs2CO3 (0.6 mmol), Pd cat 1 (0.1 mol%) and solvent (2 mL), 120 °C. b Catalyst loading (0.05 mol%). c Reaction temp = 80 °C.
1 No catalyst DMF Cs2CO3 NC
2 1 DMF No base NC
3b 1 DMF Cs2CO3 83
4 1 DMF Cs 2 CO 3 99
5 1 DMF NaOH 74
6 1 DMF Na2CO3 86
7 1 DMF KOtBu 73
8 1 Toluene Cs2CO3 54
9 1 DMSO Cs2CO3 69
10 1 THF Cs2CO3 61
11c 1 DMF Cs2CO3 58


Under these conditions, a wide range of activated aryl chlorides coupled efficiently with styrene, affording the corresponding trans-stilbenes in excellent yields (90–99%). ortho-, meta-, and para-substituted aryl chlorides bearing both electron-donating and electron-withdrawing groups were well tolerated. Notably, heteroaryl chlorides such as 2-chloronaphthalene and 2-chlorothiophene also participated effectively, providing the desired products in 90% and 92% yields, respectively (Scheme 7, entries k and r).

To investigate the reaction pathway, stoichiometric reactions of complex 1 with chlorobenzene and styrene were conducted in DMF (Scheme 8). In case of A and B, even upon heating the reaction mixture to 120 °C, no visible colour change was observed. When a mixture of complex 1, Cs2CO3, and DMF was heated in a catalytic tube at 120 °C (C), a greyish suspension formed. A similar greyish suspension was observed after 6 hours in a reaction tube containing all the reactants (Fig. 6).


image file: d5dt01353d-s8.tif
Scheme 8 Mechanistic investigation of trans-stilbene formation [(A) reaction of complex 1 with chlorobenzene, (B) reaction of complex 1 with styrene, and (C) reaction of complex 1 with Cs2CO3].

image file: d5dt01353d-f6.tif
Fig. 6 (a) Catalyst 1 and Cs2CO3, (b) catalyst 1 and Cs2CO3 upon heating at 120 °C. (c) Reaction mixture after 6 h.

To confirm the nature of the catalytic process, mercury drop and CS2 poisoning test were conducted, indicating a heterogeneous mechanism likely involving Pd nanoparticles derived from complex 1. Scanning electron microscopy images (Fig. 7(a) and (b)) revealed irregularly shaped particles in residues obtained from both (a) the reaction of complex 1 with Cs2CO3 and (b) the model catalytic reaction. These residues were isolated by DMF evaporation, followed by sequential washing with methanol, CH2Cl2, water, and methanol. The Pd nanoparticles exhibited average sizes ranging from 30–80 nm. EDX elemental mapping (Fig. S103) confirmed the presence of C, P, N, O and Pd, supporting the stabilization of Pd NPs by ligand L1 and the reductive role of Cs2CO3. To verify that the nanoparticles are indeed stabilized by the parent ligand L1 and not by decomposition products, we conducted comparative FT-IR spectroscopic analysis of free ligand L1, complex 1, and the isolated Pd nanoparticles. In the IR spectrum of ligand L1, characteristic bands were observed at 3364 cm−1 (νN–H) and 1653 cm−1 (νC[double bond, length as m-dash]O), corresponding to the amide functionalities. Upon coordination to PdII in complex 1, the νN–H band disappeared and the νC[double bond, length as m-dash]O band shifted to 1601 cm−1, clearly indicating coordination of the amide nitrogen to the metal center. For the Pd nanoparticles, IR analysis revealed the presence of both νN–H (3465 cm−1) and νC[double bond, length as m-dash]O (1666 cm−1) bands, closely resembling those of the parent ligand (Fig. 8). This observation strongly supports the retention of ligand L1 on the nanoparticle surface, suggesting that L1 acts as a stabilizing agent. The absence of new or shifted functional group frequencies also suggests that no major ligand decomposition has occurred during nanoparticle formation. These IR results, in conjunction with the EDX elemental mapping, which confirms the presence of P, N, O, and Pd, provide strong evidence that the Pd nanoparticles are indeed stabilized by the intact ligand L1.


image file: d5dt01353d-f7.tif
Fig. 7 (a) SEM image of catalyst 1 and Cs2CO3. (b) SEM image of the catalytic reaction residue.

image file: d5dt01353d-f8.tif
Fig. 8 IR spectra of red (ligand L1), green (complex 1) and blue (Pd nanoparticles).

Catalyst recyclability was evaluated over five cycles. After each run, the catalyst was recovered by solvent removal, washed, dried at 100 °C, and reused. Yields remained ≥90% under optimal conditions throughout the cycles. SEM analysis after the fifth cycle (Fig. S104) showed no significant changes in particle size or aggregation, underscoring the stability and robustness of the Pd nanoparticles. Furthermore, the practicality of the catalytic system was demonstrated through a gram-scale reaction, which afforded 97% yield of product a.

A plausible catalytic mechanism for the Heck coupling reaction is proposed based on experimental evidence and supported by literature48,79–81 (Scheme 9). Catalyst 1 acts as a Pd0 reservoir, with Cs2CO3 facilitating it's in situ reduction to Pd0 nanoparticles. The Pd nanoparticles are stabilized by ligand L1 and Cs2CO3 acts as a reductant for complex 1. The catalytic cycle begins with oxidative addition of the Pd0 species into the aryl or alkenyl halide (R-X) bond, followed by alkene insertion into the resulting Pd–Ar intermediate. Subsequent syn-β-hydride elimination affords the trans-stilbene product. Finally, elimination of hydrogen halide and its neutralization by the base regenerate the active Pd0 species, completing the cycle.


image file: d5dt01353d-s9.tif
Scheme 9 A plausible mechanism for the Heck-coupling reaction between aryl chlorides and styrene derivatives.

Conclusions

In summary, we have successfully synthesized and characterized two novel sterically hindered amide-based monophosphine ligands, L1 and L2, and explored their coordination chemistry with a range of transition metals, including PdII, CuI, AgI, and RuII. Among the resulting complexes, the dimeric P,N-coordinated palladium complex 1 exhibited outstanding catalytic performance in the Heck cross-coupling of aryl chlorides with styrene derivatives. The in situ generation of stabilized Pd nanoparticles from complex 1, facilitated by the hemilabile P,N ligand and base-induced reduction, was confirmed through SEM analysis and mercury and CS2 poisoning tests, supporting a heterogeneous catalytic mechanism. This catalytic system operated efficiently under mild conditions with low catalyst loadings (0.1 mol%), delivering trans-stilbene products in high yields (90–99%) and displaying excellent substrate scope and recyclability. These results highlight the potential of sterically encumbered P,N-type ligands in stabilizing catalytically active nanoparticles and enabling practical, selective, and sustainable C–C bond-forming transformations.

Experimental section

Instrumentation

NMR spectra were recorded on Bruker FT spectrometers (Avance-400 or 500) MHz at ambient probe temperatures. 13C{1H} and 31P{1H} NMR spectra were acquired using a broad band decoupling method. The spectra were recorded in CDCl3 and DMSO-d6 solutions with TMS as an internal standard; chemical shifts of 1H and 13C{1H} NMR spectra are reported in ppm downfield from TMS. The chemical shifts of 31P{1H} NMR spectra are referred to 85% H3PO4 as an external standard. Positive values indicate downfield shifts. Mass spectra were recorded using Bruker Maxis Impact LC-q-TOF mass spectrometer. Infrared spectra were recorded on a PerkinElmer Spectrum One FT-IR spectrometer (model no. 73465) in KBr disk. GC-MS analyses were performed on an Agilent 7890A GC system with an FID detector using a J & W DB-1 column (10 m, 0.1 mm ID). The melting points of all compounds were determined on a Veego melting point apparatus and are uncorrected. Analytical TLC was performed on a Merck 60F254 silica gel plate (0.25 mm thickness), and column chromatography was performed on Merck (100–200 MESH). The morphological studies of the thin films were conducted with a field emission scanning electron microscope (FE-SEM, JSM-7600F).

General procedure for the Mizoroki–Heck coupling reaction

The reactions were performed in a closed vessel containing a mixture of aryl halide (1 equiv.), styrene (1.2 equiv.), Cs2CO3 (1.2 equiv.), catalyst 1 (0.1 mol%) and 2 mL DMF. The reaction vessel was placed into an oil bath and heated at 120 °C. After completion of the reaction, the crude reaction mixture was treated with water (20 mL) and ethyl acetate (20 mL). The organic layer was washed with 2 × 10 mL H2O, dried over Na2SO4. The solvent was removed under reduced pressure, and the resulting crude product was purified by column chromatography on neutral or basic alumina.
Synthesis of {(o-PPh2)C6H4C(O)N(H)(C12H18)} (L1). To a thick-walled seal tube containing a magnetic stir bar were added 2-bromo-N-(2,6-diisopropylphenyl)benzamide (1 g, 2.775 mmol), Pd(PPh3)4 (0.192 g, 0.166 mmol, 6 mol%), K2CO3 (0.421 g, 3.053 mmol, 1.1 equiv.), toluene (10 mL), and PPh2H (0.671 g, 3.608 mmol, 1.3 equiv.). The tube was sealed and heated to 150 °C for 24 h with vigorous stirring. After 24 h, the reaction mixture was cooled, diluted with dichloromethane (40 mL), and washed with distilled water (3 × 30 mL). The organic layer was dried over anhydrous Na2SO4 and filtered, and the solution was concentrated under vacuum to give L1 a white colour solid, which was purified by flash chromatography over silica, eluting with 3[thin space (1/6-em)]:[thin space (1/6-em)]7 ethyl acetate/pet ether. The monophosphine L1 was characterized by multinuclear NMR spectroscopy and mass spectroscopy. Yield 75% (0.950 g). Mp: 158 °C. 1H NMR (400 MHz, CDCl3) δ 7.95 (ddd, J = 7.6, 3.8, 1.3 Hz, 1H), 7.62–7.55 (m, 2H), 7.52–7.40 (m, 7H), 7.39–7.33 (m, 5H), 7.25 (d, J = 7.7 Hz, 2H), 7.14 (dt, J = 7.7, 2.5 Hz, 1H), 3.25 (hept, J = 6.8 Hz, 2H), 1.22 (d, J = 6.9 Hz, 12H). 31P{1H} NMR (162 MHz, CDCl3) δ −11.4(s). 13C{1H} NMR (101 MHz, CDCl3) δ 168.5(s) 146.5(s), 136.8(d, J = 11 Hz), 135.2(s), 134.8(s) 134(s), 133.8(s), 131(s), 130.7(s), 129.4(s), 129(s), 128.9(s), 128.8(d, J = 7 Hz) 128.5(s), 123.6(s), 28.9(s), 23.9(s). HRMS (ESI), m/z: calcd for C31H33N1P1O1 [M + H]+: 466.2294; found 466.2272. Anal. calcd for C31H32NOP: C, 79.97; H, 6.93; N, 3.01. Found: C, 79.94; H, 6.90; N, 3.03. FT-IR (KBr disk, cm−1): 3364 (νNH) s, 2962 s, 2865 w, 1653 s (νCO), 1515 s, 1301 m, 746 s, 698 s.
Synthesis of [{(PdCl)2}{{(o-PPh2)C6H4}C(O)N(C12H18)}22-P,N] (1). Ligand L1 (0.05 g, 0.107 mmol) in 10 mL dichloromethane was added to a solution of [PdCOD(Cl)2] (0.036 g, 0.107 mmol) in 10 mL dichloromethane. The reaction mixture was allowed to stir at room temperature for 6 h. The solvent was completely removed under reduced pressure to afford complex 1 as yellow solid. The resulting solid was washed with petroleum ether (2 × 20 mL) to afford analytically pure complex 1. Yield: 83% (0.108 g). Mp: 195 °C. 1H NMR (400 MHz, CDCl3) δ 8.67 (s, 2H), 7.89 (s, 4H), 7.73 (d, J = 7.6 Hz, 3H), 7.48 (t, J = 7.7 Hz, 4H), 7.37 (t, J = 7.8 Hz, 3H), 7.25 (d, J = 8.2 Hz, 8H), 7.16 (d, J = 7.6 Hz, 3H), 6.91 (s, 5H), 6.62 (s, 2H), 2.66 (s, 4H), 0.86 (s, 24H). 31P{1H} NMR (162 MHz, CDCl3) δ 16.6(s). HRMS (ESI), m/z: calcd for C31H31ClNOPPd [M + H]+: 606.0939; found 608.0937. Anal. calcd for C62H62Cl2N2O2P2Pd2: C, 61.40; H, 5.15; N, 2.31. Found: C, 61.38; H, 5.59; N, 2.45. FT-IR (KBr disk, cm−1): 3070, 2959 m, 2853 w, 1540 m (νCO), 1462 m, 1323 s, 1102 s.
Synthesis of [{(CH3C(O)O)(Pd)2}{{(o-PPh2)C6H4}C(O)N(C12H18)}22-P,N] (2). Ligand L1 (0.05 g, 0.107 mmol) in 10 mL dichloromethane was added to a solution of Pd(OAc)2 (0.024 g, 0.107 mmol) in 10 mL dichloromethane. The reaction mixture was allowed to stir at room temperature for 6 h. The solvent was completely removed under reduced pressure to afford complex 2 as yellow solid. The resulting solid was washed with petroleum ether (2 × 20 mL) to afford analytically pure complex 2. Yield: 85% (0.114 g). Mp: 182–184 °C. 1H NMR (400 MHz, CDCl3) δ 8.46–8.43 (m, 1H), 7.67–7.43 (m, 24H), 7.33 (d, J = 7.6 Hz, 2H), 7.19 (d, J = 7.8 Hz, 2H), 7.07 (d, J = 7.6 Hz, 3H), 6.78 (dd, J = 12.3, 7.6 Hz, 2H), 3.08–2.90 (m, 4H), 1.16–0.73 (m, 30H). 13C{1H} NMR (101 MHz, CDCl3) δ 181.3(s), 164(s), 146.4(s), 145(s), 134.1(d, J = 12 Hz) 133.8(s), 132.9(d, J = 9 Hz), 132.5(s), 131.4(s), 129.8(d, J = 9 Hz), 129.4(d, J = 12 Hz), 126.9(s), 126.60(s), 126(s), 123.5(s), 122.6(s), 29(s), 24.5(s), 24.2(s). 31P{1H} NMR (202 MHz, CDCl3) δ 26.3(s). HRMS (ESI), m/z: calcd for C33H35N1P1O3Pd1 [M + H]+: 630.1384; found 630.1418. Anal. calcd for C66H68N2O6P2Pd2: C, 62.91; H, 5.44; N, 2.22. Found: C, 61.38; H, 5.19; N, 2.33. FT-IR (KBr disk, cm−1): 2961 s, 2865 w, 1549 m (νCO), 1438 w, 1263 s, 1099 s, 805 s.
Synthesis of [{(CuI)2}{{(o-PPh2)C6H4}C(O)NH(C12H18)}22-P,N] (3). CuI (0.005 g, 0.026 mmol) was suspended in acetonitrile (10 mL) and a solution of L1 (0.012 g, 0.026 mmol) in dichloromethane was added. The solution was stirred for 4 h producing a clear solution. The solution was filtered, and the solvent was removed under vacuum and residue obtained was washed with pet ether and dried to give 3 as a yellow solid. Crystals suitable for X-ray structure determination were grown by slow diffusion of pet ether into a saturated solution of the material in dichloromethane giving yellow crystals. Yield: 83% (0.116 g). Mp: 220–222 °C. 1H NMR (400 MHz, CDCl3) δ 7.86 (d, J = 7.3 Hz, 1H), 7.64–7.44 (m, 6H), 7.40–7.32 (m, 3H), 7.27–7.19 (m, 5H), 7.08 (d, J = 7.7 Hz, 2H), 6.87 (t, J = 7.9 Hz, 1H), 2.78 (s, 2H), 0.95 (d, J = 6.8 Hz, 12H). 13C{1H} NMR (101 MHz, CDCl3) δ 167.7(s), 146.6(s), 146.2(s), 134.4(d, J = 16 Hz), 132.7(s), 132.6(s), 132.5(s), 130.2(s), 130(s), 129.1(s), 128.9(d, J = 9 Hz), 128.4(s), 127.9(s), 123.6(s), 123.4(s), 28.8(s), 24.1(s). 31P{1H} NMR (162 MHz, CDCl3) δ −6.1(s). HRMS (ESI), m/z: calcd for C31H32N1P1O1Cu1 [M − I]+: 528.1512; found 528.1513. Anal. calcd for C62H64Cu2I2N2O2P2: C, 56.75; H, 4.92; N, 2.14. Found: C, 56.79; H, 4.96; N, 2.17. FT-IR (KBr disk, cm−1): 3053 (νNH) s, 2963 s, 2870 w, 1615 s (νCO), 1508 s, 1097 w, 748 s.
Synthesis of [{(AgBr)2}{{(o-PPh2)C6H4}C(O)N(H)(C12H18)}21-P] (4). Ligand L1 (0.05 g, 0.107 mmol) in 10 mL dichloromethane was added to a solution of AgBr (0.020, 0.107 mmol) in 10 mL dichloromethane. The reaction mixture was allowed to stir at room temperature for 6 h. The solvent was completely removed under reduced pressure to afford complex 4 as white solid. The resulting solid was washed with petroleum ether (2 × 20 mL) to afford analytically pure complex 4. Yield: 86% (0.120 g). Mp: 237–239 °C. 1H NMR (400 MHz, CDCl3) δ 7.85 (s, 1H), 7.52–7.09 (m, 17H), 7.01 (d, J = 10.1 Hz, 1H), 3.08 (s, 2H), 1.10 (s, 12H). 13C{1H} NMR (126 MHz, CDCl3) δ 168.3(s), 146.5(s), 134.9(s), 134(s), 133.9(s), 131(s), 130.8(s), 129.6(s), 129.3(s), 128.9(s), 128.9(s), 128.8(s), 128.5(s), 123.7(s), 28.9(s), 23.9(s). 31P{1H} NMR (162 MHz, CDCl3) δ −8.8(s). HRMS (ESI), m/z: calcd for C31H32N1P1O1Ag1 [M − Br]+: 572.1267; found 572.1266. Anal. calcd for C62H64Ag2Br2N2O2P2: C, 56.99; H, 4.94; N, 2.14. Found: C, 56.95; H, 4.92; N, 2.11. FT-IR (KBr disk, cm−1): 3318 (νNH) s, 2964 s, 2869 w, 1652 s (νCO), 1506 s, 1260 s, 1100 m, 800 s, 746 s, 696 s.
Synthesis of [{(AgBF4)2}{{(o-PPh2)C6H4}C(O)N(H)(C12H18)}23-P,O,C] (5). Ligand L1 (0.05 g, 0.107 mmol) in 10 mL dichloromethane was added to a solution of AgBF4 (0.020, 0.107 mmol) in 10 mL dichloromethane. The reaction mixture was allowed to stir at room temperature for 6 h. The solvent was completely removed under reduced pressure to afford complex 5 as white solid. The resulting solid was washed with petroleum ether (2 × 20 mL) to afford analytically pure complex 5. Yield: 81% (0.114 g). Mp: 236–238 °C. 1H NMR (400 MHz, DMSO) δ 9.61 (s, 2H), 7.87 (s, 2H), 7.78 (s, 1H), 7.59–7.26 (m, 3H), 7.34–6.65 (m, 22H), 6.65 (d, J = 7.7 Hz, 3H), 6.38 (s, 2H), 6.20 (s, 1H), 2.28–2.09 (m, 4H), 1.11 to −0.31 (m, 24H). 31P{1H} NMR (162 MHz, DMSO) δ 10.9 (d, J = 758.3 Hz). 13C{1H} NMR (101 MHz, DMSO) δ 167.7(s), 146.6(s), 139.6(s), 134.2(d, J = 17 Hz), 133.2(s), 132.9(s), 132.4(d, J = 19 Hz), 131.4(d, J = 19 Hz), 130.1(s), 129.8(s), 129.2(s), 128.6(s), 128(s), 123.5(s), 79.6(s), 28.6(s), 23.5(s). HRMS (ESI), m/z: calcd for C31H32N1P1O1Ag1 [M]+: 572.1267; found 572.1267. Anal. calcd for C62H64Ag2B2F8N2O2P2: C, 56.39; H, 4.88; N, 2.12. Found: C, 56.42; H, 4.84; N, 2.15. FT-IR (KBr disk, cm−1): 3284 (νNH) s, 2971 s, 2872 w, 1609 s (νCO), 1519 s, 1310 w, 1087 s, 749 s, 695 s.
Synthesis of [{(AgClO4)2}{{(o-PPh2)C6H4}C(O)N(H)(C12H18)}23-P,O,C] (6). Ligand L1 (0.05 g, 0.107 mmol) in 10 mL dichloromethane was added to a solution of AgClO4 (0.022, 0.107 mmol) in 10 mL dichloromethane. The reaction mixture was allowed to stir at room temperature for 6 h. The solvent was completely removed under reduced pressure to afford complex 6 as white solid. The resulting solid was washed with petroleum ether (2 × 20 mL) to afford analytically pure complex 6. Yield: 82% (0.118 g). Mp: 224–226 °C. 1H NMR (400 MHz, DMSO) δ 10.01 (s, 2H), 8.32 (d, J = 7.7 Hz, 2H), 7.77 (t, J = 7.6 Hz, 4H), 7.60 (t, J = 7.7 Hz, 4H), 7.35 (s, 14H), 7.20 (t, J = 7.7 Hz, 4H), 7.06 (d, J = 7.7 Hz, 6H), 2.58–2.53 (m, 4H), 0.88–0.73 (m, 24H). 31P{1H} NMR (162 MHz, CDCl3) δ 10.9 (d, 1JPAg 753.3). 13C{1H} NMR (101 MHz, DMSO) δ 166.9(s), 146(s), 135.6(s), 133.8(s), 133.6(s), 133(s), 132.1(s), 131.8(s), 130.8(s), 130.6(s), 129.5(s), 129.1(s), 127.9(s), 122.9(s), 79.2(s), 28(s), 22.9(s). HRMS (ESI), m/z: calcd for C31H32N1P1O1Ag1 [M]+: 572.1261; found 572.1267. Anal. calcd for C62H64Ag2Cl2N2O10P2: C, 55.33; H, 4.79; N, 2.08. Found: C, 55.37; H, 4.34; N, 2.05. FT-IR (KBr disk, cm−1): 3248 (νNH) s, 2963 s, 2866 w, 1609 s (νCO), 1516 s, 1310 w, 1084 s, 917 s, 748 s, 697 s.
Synthesis of {(o-PPh2)C6H4C(O)N(H)(C33H28)} (L2). To a thick-walled seal tube containing a magnetic stir bar were added 2-bromo-N-(2,6-dibenzhydryl-4-methylphenyl)benzamide (1 g, 1.606 mmol), Pd(PPh3)4 (0.111 g, 0.096 mmol, 6 mol%), K2CO3 (0.244 g, 1.766 mmol, 1.1 equiv.), toluene (10 mL), and PPh2H (0.388 g, 2.08 mmol, 1.3 equiv.). The tube was sealed and heated to 150 °C for 24 h with vigorous stirring. After 24 h, the reaction mixture was cooled, diluted with dichloromethane (40 mL), and washed with distilled water (3 × 30 mL). The organic layer was dried over anhydrous Na2SO4 and filtered, and the solution was concentrated under vacuum to give L2 a white colour solid, which was purified by flash chromatography over silica, eluting with 3[thin space (1/6-em)]:[thin space (1/6-em)]7 ethyl acetate/pet ether. The monophosphine L2 was characterized by multinuclear NMR spectroscopy and mass spectroscopy. Yield 82% (0.950 g). Mp: 205 °C. 1H NMR (400 MHz, CDCl3) δ 7.35–7.14 (m, 32H), 6.95 (dd, J = 7.4, 3.6 Hz, 1H), 6.57 (d, J = 2.8 Hz, 3H), 6.33 (d, J = 3.1 Hz, 1H), 5.99 (s, 2H), 2.10 (s, 3H). 13C{1H} NMR (101 MHz, CDCl3) δ 168.3(s), 143.6(s), 142.1(s), 137.6(s), 137.3(s), 134.9(s), 133.8(s), 133.6(s), 130.9(s), 130.1(s), 129.6(s), 128.6–128.3(m), 126.5(s), 126.3(s), 51.9(s), 21.7(s). 31P{1H} NMR (162 MHz, CDCl3) δ −10.6(s). HRMS (ESI), m/z: calcd for C52H43N1P1O1 [M + H]+: 728.3077; found 728.3073. Anal. calcd for C52H42NOP: C, 85.81; H, 5.82; N, 1.92. Found: C, 85.84; H, 5.84; N, 1.93. FT-IR (KBr disk, cm−1): 3401 (νNH) s, 3061 m, 3026 w, 1684 s (νCO), 1482 s, 1031 m, 753 s.
Synthesis of {(o-OPPh2)C6H4C(O)N(H)(C33H28)} (7). H2O2 (1 mL, 3.089 mmol, 30% H2O2) was added to a solution of L2 (0.050 g, 0.107 mmol) in THF (30 mL) and stirred at room temperature for 12 h. After removing the solvents under reduced pressure, the sticky oil obtained was washed with petroleum ether (2 × 20 mL) to give analytically pure compound 7 as a white solid. Single-crystals of 7 suitable for X-ray analysis were obtained by slow diffusion of petroleum ether into the dichloromethane solution of 7. Yield: 80% (0.064 g). 1H NMR (400 MHz, CDCl3) δ 7.31–7.17 (m, 25H), 7.05 (d, J = 5.1 Hz, 8H), 6.97–6.93 (m, 1H), 6.57 (s, 2H), 6.34 (s, 1H), 5.98 (s, 2H), 2.11 (s, 3H). 13C{1H} NMR (101 MHz, CDCl3) δ 168.2(s), 143.6(s), 142.1(s), 137.7(d, J = 12 Hz) 137.3(s), 134.9(s), 133.8(s), 133.6(s), 130.9(s), 130.1(s), 129.6(s), 128.6(m), 126.5(d, J = 5 Hz), 126.3(s), 51.9(s), 21.6(s). 31P{1H} NMR (162 MHz, CDCl3) δ 32.8(s). HRMS (ESI), m/z: calcd for C52H43N1P1O2 [M + H]+: 744.3028; found 744.3027. Anal. calcd for C52H42NO2P: C, 83.96; H, 5.69; N, 1.88. Found: C, 83.99; H, 5.71; N, 1.86. FT-IR (KBr disk, cm−1): 3404 (νNH) w, 3041 w, 1679 s (νCO), 1462 m, 1264 m, 1094 w, 760 s.
Synthesis of [{(CuI)2}{{(o-PPh2)C6H4}C(O)NH(C33H28)}21-P] (8). CuI (0.005 g, 0.026 mmol) was suspended in acetonitrile (10 mL) and a solution of L2 (0.018 g, 0.026 mmol) in dichloromethane was added. The solution was stirred for 4 h producing a clear solution. The solution was filtered, and the solvent was removed under vacuum and residue obtained was washed with pet ether and dried to give 8 as a yellow solid. Crystals suitable for X-ray structure determination were grown by slow diffusion of pet ether into a saturated solution of the material in dichloromethane giving yellow crystals. Yield: 81% (0.159 g). Mp: 228–230 °C. HRMS (ESI), m/z: calcd for C52H42N1P1O1Cu1 [M − I]+: 790.2295; found 790.2258. Anal. calcd for C104H84Cu2I2N2O2P2: C, 68.01; H, 4.61; N, 1.53. Found: C, 68.05; H, 4.63; N, 1.56. FT-IR (KBr disk, cm−1): 3435 (νNH) s, 3056 w, 3023 w, 1662 s (νCO), 1493 m, 1284 w, 1240 w, 744 m, 702 s. 31P{1H} NMR (162 MHz, CDCl3) δ −18.4(s).
Synthesis of [{(Ru(p-cymene)Cl)}{{(o-PPh2)C6H4}C(O)NH(C33H28)}-κ2-P,O](PF6) (9). Ligand L2 (0.05 g, 0.107 mmol) and NH4PF6 (0.017 g, 0.107 mmol) in 10 mL methanol was added to a solution of [Ru(p-cymene)Cl2]2 (0.032, 0.0535 mmol) in 10 mL methanol. The reaction mixture was refluxed at 60 °C for 24 h. The reaction mixture was filtered through Celite and the solvent was completely removed under reduced pressure to afford complex 9 as orange solid. The resulting solid was washed with diethyl ether (2 × 20 mL) to afford analytically pure complex 9. Yield: 86% (0.105 g). 1H NMR (400 MHz, CDCl3) δ 9.43 (s, 1H), 8.36 (s, 1H), 7.77 (dd, J = 11.8, 7.6 Hz, 2H), 7.69–7.43 (m, 11H), 7.29 (s, 9H), 7.12 (dd, J = 7.8, 1.4 Hz, 1H), 6.85 (dd, J = 11.4, 7.6 Hz, 1H), 5.83 (d, J = 5.5 Hz, 1H), 5.59 (d, J = 5.8 Hz, 1H), 5.48 (d, J = 5.6 Hz, 1H), 5.23 (d, J = 5.6 Hz, 1H), 1.36 (d, J = 6.6 Hz, 3H), 0.80 (d, J = 6.9 Hz, 3H), 0.74 (d, J = 6.8 Hz, 3H), 0.65 (d, J = 6.2 Hz, 3H). 31P{1H} NMR (162 MHz, CDCl3) δ 32.5, −144.3 (sept, 1JPF 714.4 Hz). 13C{1H} NMR (101 MHz, CDCl3) δ 170.8(s), 146.1(s), 145.9(s), 134.6(d, J = 11 Hz), 133.4(d, J = 9 Hz), 132.9–132.3(m), 131.9(s), 131.3(s), 131.2(s), 130.7(s), 129.9(d, J = 9 Hz), 129.3(d, J = 11 Hz), 129(s), 128.9(s), 128.7(s), 128.5(s), 128.1(s), 126.9(s), 123.4(d, J = 3 Hz), 104.3(s), 100(s), 88.8(s), 87.8(s), 86.9(s), 86.8(s), 29.1(s), 28.8(s), 24.6(s), 23.2(s), 20.2(s), 16.9(s). Anal. calcd for C62H56ClF6NOP2Ru: C, 65.12; H, 4.94; N, 1.22. Found: C, 65.16; H, 4.38; N, 1.32. FT-IR (KBr disk, cm−1): 3397 (νNH) w, 3070 w, 2493 w, 1597 m (νCO), 1519 w, 1442 w, 1323 w, 1259 w, 1102 w, 838 s, 704 s.

Author contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included as part of the ESI.

ESI contents: crystal structure determination of compounds; NMR and HRMS spectra of complexes; controlled experiments of reaction mechanism; NMR and mass spectra of catalytic products.

Acknowledgements

MSB thank Indian Institute of Technology Bombay for supporting this work through Research Development Fund (RDF). We are thankful to the Department of Chemistry, IIT Bombay, for instrumentation facilities, as well as spectral and analytical data. GS acknowledges the financial support from IITB and KCD thanks UGC for the fellowship.

References

  1. Y. Canac, Chem. Rec., 2023, 23, e202300187 CrossRef CAS PubMed.
  2. M. R. Tiddens and M.-E. Moret, in Metal-Ligand Co-operativity: Catalysis and the Pincer-Metal Platform, ed. G. van Koten, K. Kirchner and M.-E. Moret, Springer International Publishing, Cham, 2021, pp. 25–69 Search PubMed.
  3. S. Maggini, Coord. Chem. Rev., 2009, 253, 1793–1832 CrossRef CAS.
  4. K. N. Gavrilov and A. I. Polosukhin, Russ. Chem. Rev., 2000, 69, 661–682 CrossRef CAS.
  5. P. J. Guiry and C. P. Saunders, Adv. Synth. Catal., 2004, 346, 497–537 CrossRef CAS.
  6. L. Radhakrishna, B. S. Kote, H. S. Kunchur, M. K. Pandey, D. Mondal and M. S. Balakrishna, Dalton Trans., 2022, 51, 5480–5493 RSC.
  7. M. P. Carroll and P. J. Guiry, Chem. Soc. Rev., 2014, 43, 819–833 RSC.
  8. L. Alig, M. Fritz and S. Schneider, Chem. Rev., 2019, 119, 2681–2751 CrossRef CAS PubMed.
  9. J. Bae and E. J. Cho, ACS Catal., 2023, 13, 13540–13560 CrossRef CAS.
  10. H. S. Kunchur, L. Radhakrishna, M. K. Pandey and M. S. Balakrishna, Chem. Commun., 2021, 57, 4835–4838 RSC.
  11. H. S. Kunchur and M. S. Balakrishna, Inorg. Chem., 2022, 61, 857–868 CrossRef CAS PubMed.
  12. B. Michelet, D. Lebœuf, C. Bour, K. Škoch, F. Horký, P. Štěpnička and V. Gandon, ChemPlusChem, 2017, 82, 442–448 CrossRef CAS PubMed.
  13. B. T. Ingoglia and S. L. Buchwald, Org. Lett., 2017, 19, 2853–2856 CrossRef CAS PubMed.
  14. B. V. Rokade and P. J. Guiry, ACS Catal., 2018, 8, 624–643 CrossRef CAS.
  15. M. K. Yılmaz, S. İnce, M. Keleş and B. Güzel, J. CO2 Util., 2020, 42, 101309 CrossRef.
  16. B. S. Kote, M. A. Mohite, D. Mondal, M. K. Pandey and M. S. Balakrishna, Eur. J. Inorg. Chem., 2023, 26, e202300291 CrossRef CAS.
  17. J. I. van der Vlugt, J. M. J. Paulusse, E. J. Zijp, J. A. Tijmensen, A. M. Mills, A. L. Spek, C. Claver and D. Vogt, Eur. J. Inorg. Chem., 2004, 2004, 4193–4201 CrossRef.
  18. A. Aghmiz, A. M. Masdeu-Bultó, C. Claver and D. Sinou, J. Mol. Catal. A: Chem., 2002, 184, 111–119 CrossRef CAS.
  19. F. Schneck, M. Finger, M. Tromp and S. Schneider, Chem. – Eur. J., 2017, 23, 33–37 CrossRef CAS PubMed.
  20. R. González-Fernández, P. Crochet and V. Cadierno, Dalton Trans., 2020, 49, 210–222 RSC.
  21. D. G. A. Verhoeven and M.-E. Moret, Dalton Trans., 2016, 45, 15762–15778 RSC.
  22. M. K. Pandey, J. T. Mague and M. S. Balakrishna, Inorg. Chem., 2018, 57, 7468–7480 CrossRef CAS PubMed.
  23. R. Gourkhede, B. Kaur, B. S. Kote and M. S. Balakrishna, Dalton Trans., 2024, 53, 10693–10703 RSC.
  24. M. R. Tiddens, B. T. Kappé, T. J. Smak, M. Lutz and M.-E. Moret, Chem. – Eur. J., 2024, 30, e202400666 CrossRef CAS PubMed.
  25. S. S. Ng, Z. Chen, O. Y. Yuen and C. M. So, Org. Biomol. Chem., 2022, 20, 1373–1378 RSC.
  26. W.-M. Dai and Y. Zhang, Tetrahedron Lett., 2005, 46, 1377–1381 CrossRef CAS.
  27. X. F. Bai, T. Song, Z. Xu, C. G. Xia, W. S. Huang and L. W. Xu, Angew. Chem., 2015, 127, 5344–5348 CrossRef.
  28. W.-M. Dai, K. K. Y. Yeung, J.-T. Liu, Y. Zhang and I. D. Williams, Org. Lett., 2002, 4, 1615–1618 CrossRef CAS PubMed.
  29. W.-M. Dai, Y. Li, Y. Zhang, C. Yue and J. Wu, Chem. – Eur. J., 2008, 14, 5538–5554 CrossRef CAS PubMed.
  30. M. H. Tse, P. Y. Choy and F. Y. Kwong, Acc. Chem. Res., 2022, 55, 3688–3705 CrossRef CAS PubMed.
  31. W. Li and J. Zhang, Acc. Chem. Res., 2024, 57, 489–513 CAS.
  32. N. W. Boaz, J. A. Ponasik, S. E. Large and S. D. Debenham, Tetrahedron: Asymmetry, 2004, 15, 2151–2154 CrossRef CAS.
  33. M. Tamizmani, R. Kankanala and C. Sivasankar, J. Organomet. Chem., 2014, 763–764, 6–13 CrossRef CAS.
  34. B. Tezcan, M. Kemal Yılmaz, G. Yakalı, M. Aygün and B. Güzel, Inorg. Chim. Acta, 2022, 543, 121155 CrossRef CAS.
  35. L. Yin and J. Liebscher, Chem. Rev., 2007, 107, 133–173 CrossRef CAS PubMed.
  36. K. A. Jantan, G. Ekart, S. McCarthy, A. J. P. White, D. C. Braddock, A. Serpe and J. D. E. T. Wilton-Ely, Catalysts, 2024, 14, 295 CrossRef CAS.
  37. E. A. Onoabedje and U. C. Okoro, Synth. Commun., 2019, 49, 2117–2146 CrossRef CAS.
  38. A. Biffis, P. Centomo, A. Del Zotto and M. Zecca, Chem. Rev., 2018, 118, 2249–2295 CrossRef CAS PubMed.
  39. N. Selander and K. J. Szabó, Chem. Rev., 2011, 111, 2048–2076 CrossRef CAS PubMed.
  40. R. Taakili, C. Barthes, C. Lepetit, C. Duhayon, D. A. Valyaev and Y. Canac, Inorg. Chem., 2021, 60, 12116–12128 CrossRef CAS PubMed.
  41. A. Balanta, C. Godard and C. Claver, Chem. Soc. Rev., 2011, 40, 4973–4985 RSC.
  42. Z. Wang, P. Xiao, B. Shen and N. He, Colloids Surf., A, 2006, 276, 116–121 CrossRef CAS.
  43. A. Bej, K. Ghosh, A. Sarkar and D. W. Knight, RSC Adv., 2016, 6, 11446–11453 RSC.
  44. K. Hong, M. Sajjadi, J. M. Suh, K. Zhang, M. Nasrollahzadeh, H. W. Jang, R. S. Varma and M. Shokouhimehr, ACS Appl. Nano Mater., 2020, 3, 2070–2103 CrossRef CAS.
  45. J. Wang, W. S. Cheon, J.-Y. Lee, W. Yan, S. Jung, H. W. Jang and M. Shokouhimehr, Dalton Trans., 2023, 52, 3567–3574 RSC.
  46. S. Ince, Ö. Öner, M. K. Yılmaz, M. Keleş and B. Güzel, Inorg. Chem., 2023, 62, 4637–4647 CrossRef CAS PubMed.
  47. M. Tamura and H. Fujihara, J. Am. Chem. Soc., 2003, 125, 15742–15743 CrossRef CAS PubMed.
  48. S. Sheokand and M. S. Balakrishna, Inorg. Chem., 2023, 62, 12317–12328 CrossRef CAS PubMed.
  49. S. McCarthy, D. C. Braddock and J. D. E. T. Wilton-Ely, Coord. Chem. Rev., 2021, 442, 213925 CrossRef CAS.
  50. M. Alaqarbeh, S. F. Adil, T. Ghrear, M. Khan, M. Bouachrine and A. Al-Warthan, Catalysts, 2023, 13, 1343 CrossRef CAS.
  51. C. Deraedt and D. Astruc, Acc. Chem. Res., 2014, 47, 494–503 CrossRef CAS PubMed.
  52. F. Ulusal, E. Erünal and B. Güzel, J. Nanopart. Res., 2018, 20, 219 CrossRef.
  53. P. J. Anju, M. Neetha and G. Anilkumar, ChemistrySelect, 2022, 7, e202103564 CrossRef CAS.
  54. R. F. Heck and J. Nolley Jr, J. Org. Chem., 1972, 37, 2320–2322 CrossRef CAS.
  55. K. S. Madden, S. David, J. P. Knowles and A. Whiting, Chem. Commun., 2015, 51, 11409–11412 RSC.
  56. S. Mohanty and M. S. Balakrishna, J. Chem. Sci., 2010, 122, 137–142 CrossRef CAS.
  57. S. Mohanty, D. Suresh, M. S. Balakrishna and J. T. Mague, Tetrahedron, 2008, 64, 240–247 CrossRef CAS.
  58. D. Mondal and M. S. Balakrishna, Eur. J. Inorg. Chem., 2020, 2020, 2392–2402 CrossRef CAS.
  59. I. P. Beletskaya and A. V. Cheprakov, Chem. Rev., 2000, 100, 3009–3066 CrossRef CAS PubMed.
  60. O. Blacque and C. M. Frech, Chem. – Eur. J., 2010, 16, 1521–1531 CrossRef CAS PubMed.
  61. M. Oberholzer and C. M. Frech, Green Chem., 2013, 15, 1678–1686 RSC.
  62. D. Mc Cartney and P. J. Guiry, Chem. Soc. Rev., 2011, 40, 5122–5150 RSC.
  63. M. Oberholzer and C. M. Frech, J. Visualized Exp., 2014, e51444,  DOI:10.3791/51444.
  64. K. Hirabayashi, Y. Nishihara, A. Mori and T. Hiyama, Tetrahedron Lett., 1998, 39, 7893–7896 CrossRef CAS.
  65. K. Kikukawa and T. Matsuda, Chem. Lett., 1977, 6, 159–162 CrossRef.
  66. Y. Hou, J. Ma, H. Yang, E. A. Anderson, A. Whiting and N. Wu, Chem. Commun., 2019, 55, 3733–3736 RSC.
  67. M. Beller, H. Fischer, W. A. Herrmann, K. Öfele and C. Brossmer, Angew. Chem., Int. Ed. Engl., 1995, 34, 1848–1849 CrossRef CAS.
  68. S. Jagtap, Catalysts, 2017, 7, 267 CrossRef.
  69. V. W. Bhoyare, E. D. Sosa Carrizo, C. C. Chintawar, V. Gandon and N. T. Patil, J. Am. Chem. Soc., 2023, 145, 8810–8816 CrossRef CAS PubMed.
  70. S. Mukhopadhyay, G. Rothenberg, A. Joshi, M. Baidossi and Y. Sasson, Adv. Synth. Catal., 2002, 344, 348–354 CrossRef CAS.
  71. H.-J. Xu, Y.-Q. Zhao and X.-F. Zhou, J. Org. Chem., 2011, 76, 8036–8041 CrossRef CAS PubMed.
  72. W. A. Herrmann, C. Brossmer, C.-P. Reisinger, T. H. Riermeier, K. Öfele and M. Beller, Chem. – Eur. J., 1997, 3, 1357–1364 CrossRef CAS.
  73. W. Chen, I. Shaikh, F. Ahmed, S. Karkoub, M. AlRawashdeh, H. Zhou and S. Madrahimov, ChemistryOpen, 2024, 13, e202300249 CrossRef CAS PubMed.
  74. N. Kvasovs, V. Iziumchenko, V. Palchykov and V. Gevorgyan, ACS Catal., 2021, 11, 3749–3754 CrossRef CAS PubMed.
  75. S. Lerch, S. Fritsch and T. Strassner, Eur. J. Org. Chem., 2022, e202200008 CrossRef CAS.
  76. G. Sabharwal, K. C. Dwivedi, C. Das, T. R. K. Rana, A. Dutta, G. Rajaraman and M. S. Balakrishna, J. Catal., 2024, 440, 115825 CrossRef CAS.
  77. K. C. Dwivedi, G. Sabharwal, B. S. Kote and M. S. Balakrishna, Dalton Trans., 2024, 53, 18321–18329 RSC.
  78. K. C. Dwivedi, G. Sabharwal, A. Pandey and M. S. Balakrishna, Inorg. Chem., 2025, 64, 10989–11000 CrossRef CAS PubMed.
  79. J. P. Knowles and A. Whiting, Org. Biomol. Chem., 2007, 5, 31–44 RSC.
  80. V. P. W. Böhm and W. A. Herrmann, Chem. – Eur. J., 2001, 7, 4191–4197 CrossRef.
  81. A. M. Trzeciak and J. J. Ziółkowski, Coord. Chem. Rev., 2005, 249, 2308–2322 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. CCDC 2456488–2456497. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5dt01353d

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.