Open Access Article
Amin
Abdolrahimi
,
Philipp
Woite
,
Konrad
Kretschmar
,
Michael
Roemelt
*,
Thomas
Braun
* and
Ouchan
He
Department of Chemistry, Humboldt-Universität zu Berlin, Brook-Taylor Str. 2, 12489 Berlin, Germany. E-mail: thomas.braun@cms.hu-berlin.de
First published on 15th September 2025
The paper introduces bimetallic Ir/Pd complexes as catalysts to initiate an N-methylation coupled with an aminocarbonylation in a one-pot approach, in order to advance the field of carboxyamide synthesis. The iridium center initiates the reaction by selectively facilitating the N-methylation through amination, while the palladium center, in a complementary role, drives the carbonylation step. This bimetallic synergy not only streamlines the reaction sequence but also surpasses the efficiency and selectivity of monometallic Ir or Pd catalysts. Mechanistic studies suggest the presence of catalytically active hydride species within the N-methylation cycle, which were characterized experimentally and via quantum chemical calculations. The developed synthetic routes offer a sustainable, cost-effective, scalable and also unprecedented preparation method, with the bimetallic catalysts being robust and versatile.
Bimetallic catalysts, particularly heterobimetallic ones, are promising for sequential catalysis, where two different metal centers synergistically catalyze multiple reaction steps. This can led to improved or unique reactivities and selectivities compared to homobimetallic or monometallic catalysts.10–32 While processes enhancing the activity of heterobimetallic complexes compared to their monometallic counterparts are well-established, the one-pot activity for sequential reaction pathways remains limited.33–38 An array of late–late bimetallic transition metal complexes has been synthesized to serve as unique catalysts in diverse C–N/C–C bond coupling reactions.39–43 Notably, bimetallic Pd(I) and Pd(II) complexes exhibit activity in aminocarbonylation of primary amines and Buchwald–Hartwig aminations.44 A heterobimetallic catalyst featuring Ir/Pd centers and a 1,2,4-trimethyltriazolyldiylidene ligand (ditz) was employed to catalyze tandem imine formation from corresponding nitroarenes.45 Furthermore, three sets of Ir/Pd complexes with the mentioned bridging ligand were synthesized and utilized in tandem reactions involving various transformations of halo-acetophenones. These transformations encompass (i) dehalogenation and transfer hydrogenation, (ii) Suzuki coupling and transfer hydrogenation, and (iii) Suzuki coupling and α-alkylation with primary alcohols.46
In this paper we report on the synthesis of Ir/Pd bimetallic complexes and their application in a one-pot process for mono-N-methylations of primary amines under air, coupled with aminocarbonylations of the resulting secondary amine under atmospheric CO pressure. Model studies gave mechanistic insights by identifying catalytically relevant hydride species, which were also characterized by means of quantum chemical calculations (Fig. 1).
![]() | ||
| Fig. 1 (a) Transition metal-catalyzed N-alkylation. (b) Pd-catalyzed aminocarbonylation. (c) Sequential one-pot N-alkylation followed by aminocarbonylation of primary amines. | ||
Routes to bimetallic Ir/Pd complexes bearing imidazole or pyridine based bridging phosphine ligands were developed. The bimetallic compounds [IrPd(Cl)3(CO)(PiPr2Im)2] (Im = imidazole) (4), [IrPd(Cl)3(CO)(PiPr2ImMe)2] (ImMe = 1-methyl-1H-imidazole) (5) and [IrPd(Cl)3(CO)(PPh2Py)2] (Py = pyridine) (6) were synthesized by treatment of the corresponding mononuclear Ir complexes trans-[Ir(Cl)(CO)(PiPr2Im)2] (1), trans-[Ir(Cl)(CO)(PiPr2ImMe)2] (2) and trans-[Ir(Cl)(CO)(PPh2Py)2] (3) with [Pd(Cl)2(COD)] (Scheme 1). The complexes 3 and 6 (Scheme 1) have been reported before, but 6 was synthesized via an alternative route.47 Other binuclear complexes featuring 2-(diphenylphosphino)pyridine (Ph2PPy) as a bridging ligand with iridium centers have also been described.49–51 The bimetallic Ir/Pd complexes were characterized by 1H NMR, 31P{1H} NMR, and IR spectroscopy (see SI). Notably, the IR spectrum of 4 displays a very broad absorption band at
= 3091 cm−1 for the NH unit, which is shifted to lower frequencies compared to the NH absorption frequency of 1 at 3343 cm−1. The NMR and IR data suggest the presence of hydrogen bonding.52,53 Strong absorption bands for the terminal-bound CO were detected between 1950 and 2016 cm−1 for all three complexes 4–6 in the ATR-IR spectra (4 (
= 2000 cm−1), 5 (
= 1950 cm−1) and 6 (
= 2013 cm−1).47Table 1 shows these IR frequencies together with their corresponding 31P NMR data.
![]() | ||
| Scheme 1 Routes to the complexes 4, 5, 3 and 6.47 | ||
| Compound | CO stretching frequencies ( ) |
31P NMR chemical shift (ppm) |
|---|---|---|
| 1 | 1944 | 28.9 |
| 2 | 1942 | 25.9 |
| 3 | 1971 | 24.4 |
| 4 | 2000 | −0.76 |
| 5 | 1950 | 10.54 |
| 6 | 2013 | −12.50 |
The structure of the complex [IrPd(Cl)3(CO)(PiPr2Im)2] (4) in the solid state was determined by X-ray crystal structure analysis (Fig. 2).
The distance between Ir and Pd (2.6259(4) Å) is in the typical range of Ir–Pd metal bonds of similar complexes, for instance 2.614(1) Å for 6 (ref. 47) and 2.694(2) Å for [IrPd(Cl)(CO)(dpmp)2][PF6]2 (ref. 48) (dpmp: bis-((diphenylphosphino)methyl)-phenylphosphine). The amine functions (NH) of the two imidazolylphosphines each form a hydrogen bond to two DMSO solvent molecules in the asymmetric unit.52,53
Efforts to get single crystals of the heterobimetallic complex 5 for X-ray analysis proved to be unsuccessful. DFT calculations using a well-tested hybrid-functional (TPSSh) were conducted to determine the structure of 5, which is illustrated in Fig. 3a. The predicted Ir–Pd bond length in complex 5 (2.601 Å) is shorter than the corresponding distances in the complexes 4 and 6.47 In the heterobimetallic complex 5, the highest occupied molecular orbital (HOMO) is predominantly distributed between the two metal centers, indicating significant metal–metal interaction (Fig. 3b, see SI for the corresponding molecular orbitals of 4 and 6).
In initial catalytic studies aniline was selectively methylated using methanol, KOtBu, and 0.5 mol% catalyst at 100 °C under air atmosphere, to yield N-methylaniline (10a) in 4 hours. The catalytic activities of the monometallic Ir complexes 1, 2 and 3 were compared to those of the bimetallic catalysts 4, 5 and 6. The performance of bimetallic catalysts is for all cases under the given conditions slightly better over the monometallic Iridium complexes (entries 2–7, Table 2). Kinetic profiles for the bimetallic catalysts 4, 5 and 6 revealed that catalyst 5 gave rise to 99% yield within 3.5 hours. Catalysts 4 and 6 showed a comparable activity, reaching similar yields at a slower rate. Note that with [Ir(Cl)(COD)]2 (ref. 70) as a catalyst, double methylation occurred (entry 1, Table 2).54,55
| Entry | Cat. | Base | T (°C) | Yieldb (%) |
|---|---|---|---|---|
| a Reaction conditions: aniline (1 mmol), methanol (0.2 ml), catalyst (0.5 mol%), base (150 mol%), at 100 °C for 4 h under air atmosphere. b Yields were determined by GC-MS with mesitylene as the internal standard. c Reacting for 6 h, the selectivites towards N-methylaniline and N,N-dimethyl-aniline were 10% and 90% respectively. d Within 3.5 h. | ||||
| 1c | [Ir(Cl)(COD)]2 | KOtBu | 100 | 14 |
| 2 | 1 | KOtBu | 100 | 83 |
| 3 | 2 | KOtBu | 100 | 85 |
| 4 | 3 | KOtBu | 100 | 83 |
| 5 | 4 | KOtBu | 100 | 96 |
| 6 | 5 | KOtBud | 100 | 99 |
| 7 | 6 | KOtBu | 100 | 93 |
| 8 | 1 + 7 | KOtBu | 100 | 53 |
| 9 | 2 + 8 | KOtBu | 100 | 74 |
| 10 | 3 + cis/trans-9 | KOtBu | 100 | 75 |
| 11 | 5 | K2CO3 | 100 | 47 |
| 12 | 5 | Cs2CO3 | 100 | 59 |
| 13 | 5 | KOtBu | 60 | 55 |
| 14 | 5 | KOtBu | 25 | 30 |
Importantly, mixtures of the monometallic Ir (1–3) and Pd (7–9)56,57 catalysts (Fig. 4) showed much lower conversions under the same conditions (entries 8–10, Table 2). Replacing KOtBu with K2CO3 or Cs2CO3 as the base resulted in significantly lower conversion when using 5 as the catalytic precursor (entries 11–12, Table 2). The transformation with 5 was also performed at 60 °C and room temperature (entries 13 and 14, Table 2). Although, the activities were lower showing 30% yield for the latter, homogeneous N-methylation at room temperature is unprecedented.58,59
The scope of N-methylation was tested with various aniline derivatives, and catalyst 5 was efficiently able to mono-N-methylate diverse aromatic amines (Scheme 2). This includes p-substituted aromatic anilines with p-Me (10b), p-OMe (10c), and various halides (10d–10f), producing excellent yields (99%). The sterically hindering o-Me (10g) and o-OMe (10h) moieties slightly reduced reactivity but still produced excellent yields for o-F (10i). Additionally, 3,5-substituted anilines showed a complete conversion (10j–10l). Other aniline derivatives, such as 3.4-dimethyl, 2,4-dimethyl, pyridine, and naphthalene (10m–10q), also achieved nearly excellent yields (94–99%) of mono-N-methylated products.
Mechanistically it can be assumed that methanol undergoes hydrogen transfer dehydrogenation to form formaldehyde by a borrowing hydrogen mechanism mediated by the Ir center.60–74 The N-methylamine product is then obtained by hydrogenation of the in situ generated imine.70,74 To gain mechanistic insights into the N-methylation catalytic cycle, we examined the reactivity of complexes 4, 5, and 6 in independent reactions. When complex 5 was treated with MeOH under the optimized catalytic conditions, it produced the unique binuclear trihydrido complex 5H3 and a dihydrido complex 5H2 in a 1.6
:
1 ratio based on the 1H NMR data (Scheme 3a). Complex 6 yielded a hydrido complex (6Ha or 6Hb, Scheme 3b), which exhibits a hydrido ligand in the cis position to the CO ligand (Scheme 3b; (1H NMR) δ = −16.58 ppm). The reaction of complex 4 with MeOH generated a dihydrido complex in low yields, which could not be characterized further.
Independent routes to generate the complexes 5H2 and 6Ha/6Hb were also developed (Scheme 4). Complex 2 generated with MeOH and KOtBu to give the mononuclear hydrido iridium complex 2H3, which further reacted with [Pd(Cl)2(COD)] to yield the bimetallic complex 5H2 (Scheme 4a). Note that a synthesis of the complex 5H3 proved unfeasible by treatment of 5H2 with HCl. Upon exposure of a solution of complex 3 to H2, the dihydrido complex 3H2 formed, as verified by X-ray crystallography (see SI). 3H2 was then treated with [Pd(Cl)2(COD)], producing in this case two isomeric monohydride complexes of 6Ha/6Hb (Scheme 4b), one of which is the one that is also produced from 6 by reaction with MeOH.
1H NMR data for 5H3 showed that all the hydrido ligands are located in a cis position to the phosphine ligands. Two signals appear as triplet of doublets and triplet of triplets signals at (δ = −11.97 and −13.07 ppm, respectively) which integrate 2
:
1 (see SI). 2JH,P coupling constants of 14.7 Hz and 19.8 Hz confirm the cis positions to phosphines. However, the latter hydride has a trans position to the CO ligand as confirmed by NMR data of the 13CO labeled isotopomer of 5H3 (see SI).47 For 5H2 two sets of triplet of doublet signals (δ = −8.91 and −20.43 ppm) with an 1
:
1 integration indicate the presence of two hydrido ligands (2JH,P coupling constants of 18.5 Hz and 13.0 Hz, respectively).
The IR spectra of 5H2 and 5H3 showed three bands each: 1947, 1982, and 2097 cm−1 for 5H2, and 2115, 1997, and 2158 cm−1 for 5H3. These can be assigned to CO and two hydrido ligands at the iridium centers;47 the data fit to the calculated IR spectra for 5H2 and 5H3 (1940/1998/2099 cm−1 and 2133/2003/2169 cm−1, see below).
Several attempts to obtain single crystals from the 5H3/5H2 mixture were unsuccessful due to their low stability. Therefore, DFT was used to analyze the intermediates. Fig. 5 shows the optimized structures for 5H2 and 5H3. Based on these structures the Ir–Pd bond lengths are 2.655 Å and 2.644 Å for 5H3 and 5H2 respectively. The Ir center in 5H3 features a distorted coordination environment resulting in two hydrido ligands that could be considered magnetically equivalent as well as a bridging hydrido ligand. A comparable configuration has also been reported for binuclear Ir/Ir and Ir/Rh complexes.25 Other possible isomers for 5H3 did not converge.
The 1H NMR spectra of the two isomers 6Ha/6Hb displayed two triplet signals at δ = −15.39 and −16.58 ppm (2JH,P = 11.5 and 15.0 Hz respectively) in a 1
:
1 ratio. 1H NMR spectra of the 13C isotopologues confirm the hydrides to be in the cis positions to the respective 13CO ligands (2JH,C = 5.2 and 3.1 Hz, respectively).47 The IR spectrum of a dichloromethane solution containing the isomers 6Ha and 6Hb revealed CO absorption bands at 2047 and 1994 cm−1,47 along with bands for the hydrido ligands at 2180 and 2192 cm−1. The relative energies of the two isomers were calculated by DFT. 6Ha, which is defined to be the complex with the hydrido ligand trans to Cl, is 3.15 kcal mol−1 more stable than 6Hb. Another hypothetical structure with the hydride in the trans position to the CO ligand 6Hc is 2.33 kcal mol−1 less stable than 6Ha. However, 6Hc was ruled out by NMR spectroscopy (see above), despite thermodynamic accessibility.
Monitoring the reaction solution for the conversion of 4-fluoroaniline to give 10d with 5 as pre-catalyst by 1H NMR and 19F NMR spectroscopy revealed the presence of the hydrido species 5H3 and 5H2, indeed (see SI). Another model reaction showed that the hydrides can act as hydrogen sources for imine hydrogenation. Thus, treatment of N-benzylideneaniline with a solution of 5H3/5H2 or 6Ha/6Hb in the presence of methanol led to the generation of the corresponding amine with 34% and 15% yield, as confirmed by GC-MS (Scheme 5a). Note that 5H2 is more reactive as at the end of the reaction only small amounts of 5H3 remained. The mononuclear species 2H3, however, was slightly less reactive under these conditions, while 3H2 showed no detectable conversion.64–69
Notably, catalytic N-methylation experiments using methanol solutions of 5H3/5H2, 6Ha/6Hb produced 10a in marginally higher yields than those reactions employing the monometallic iridium hydrides 2H3 and 3H2 (Scheme 5b). Experiments with methanol-d4 and aniline resulted in the formation of labeled secondary amine. The kinetic isotope effect (KIE) for the N-methylation of aniline was also studied on using CH3OH and CD3OD. The reactions were monitored by GC-MS with mesitylene as an internal standard and a KIE value of 2.04 was determined from the ratio of initial rates (see SI).70,74
Under the optimized conditions using the bimetallic catalysts for N-methylation (entries 5–7, Table 3), we then investigated the aminocarbonylation of N-methylaniline. Thus, reactions of 10a with 4-iodotoluene were run under CO atmosphere in the presence of a base using 4, 5, and 6 as pre-catalysts (entries 1–3, Table 3). With 1 mol% of 5 amide 11f was formed with 85% yield when the reaction was run in the presence of K2CO3 as base in toluene at 100 °C within 16 h under atmospheric CO pressure (entry 1, Table 3). 5 showed a slightly better activity than 4 or 6 (entries 1 and 3, Table 3). As the carbonylation steps are presumably mainly mediated by the Pd center,75–94 the bimetallic systems were compared with mononuclear Pd catalysts (7, 8 and 9), all showing a significant decrease in the 11f generation (entries 4–6, Table 3). Remarkably, testing mixtures of mononuclear Ir and Pd complex corresponding to the bimetallic complexes also resulted in reduced yields (entries 7 and 8, Table 3). Moreover, using K3PO4 as a base with 5 as the catalyst or DMF as a solvent also gave a reduced yield (entry 9–11, Table 3). Overall, bimetallic systems outperform mononuclear counterparts in carbonylation of secondary amines, suggesting on a cooperative effect between two metal centers. Presumably, the carbonylation process is predominantly mediated by the Pd center. A classic reaction sequence involves an oxidative addition of the aryl halide, an amido complex formation, insertion of CO into the Pd–C bond, and subsequent reductive elimination of the C–N bond to yield the amide product;79,81,83 however, the involvement of the iridium centre in this transformation remains uncertain.
| Entry | Cat. | Base | Solvent | Yieldb (%) |
|---|---|---|---|---|
| a Reaction conditions: cat. (1 mol%), amine (1 mmol), aryl iodide (1.2 mmol), base (2 eq) and CO (1 bar). b Yields were determined by GC-MS with mesitylene as the internal standard. | ||||
| 1 | 4 | K2CO3 | Toluene | 80 |
| 2 | 5 | K2CO3 | Toluene | 99 |
| 3 | 6 | K2CO3 | Toluene | 77 |
| 4 | 7 | K2CO3 | Toluene | 50 |
| 5 | 8 | K2CO3 | Toluene | 51 |
| 6 | 9 | K2CO3 | Toluene | 45 |
| 7 | 2 + 8 | K2CO3 | Toluene | 30 |
| 8 | 3 + 9 | K2CO3 | Toluene | 35 |
| 9 | 5 | K3PO4 | Toluene | 70 |
| 10 | 5 | K2CO3 | DMF | 35 |
| 11 | 5 | Et3N | DMF | 40 |
In model reactions, toluene solutions of 5H2 and 5H3 were then reacted with 4-iodobenzotrifluoride. The formation of trifluorotoluene as a hydrodehalogenated product with 35% yield was observed, whereas the 1H NMR spectra showed various hydride complexes, which were not characterized further. Repeating the reaction with 6Ha/6Hb did not produce trifluorotoluene.
The complexes 5H2 and 6Ha/6Hb were then tested in the carbonylation of 10a with 4-iodotoluene to yield successfully 11b with 62% and 43% yield, respectively (Scheme 5c). The mononuclear Pd catalyst 8 delivered a slightly lower yield (51%) than 5H2 (Scheme 5c). The Ir hydride species 2H3 and 3H2 showed no catalytic activity. Note that 5H3 is not stable in the presence of CO, and the monohydrido complex [IrPdCl2(H)(CO)(PiPr2ImMe)2] (5H) is formed under H2 release, as confirmed by the 1H NMR spectrum. However, after five days the generation of 5H2 was observed (see SI). Furthermore, upon exposure of complex 5 to 1 bar atmosphere of 13CO, the formation of 13CO-5 was observed. Note that 13CO-5 could also be generated alternatively from 13CO-2 carbonyl ligand exchange (see SI). In situ–formed 13CO-5 was then treated in a stoichiometric reaction with 4-iodobenzotrifluoride and 10a in the absence of additional CO. The generation of the 13C labelled carbonylation product 13CO-11j was observed (Scheme 5d). This indicates a partial participance of the Ir centre during the carbonylation step as it can provide the CO moiety for carbonylation.
The scope of the catalytic synthesis of carboxyamides was then explored on running consecutive one-pot N-methylation and aminocarbonylation reactions (Scheme 6). It has to be emphasized that these conversions can also be classified as assisted-relay catalysis.7 Complex 5 enabled the sequential N-methylation of aniline derivatives with MeOH and subsequent carbonylation with iodobenzene to achieve 90–99% yields for 11a–11e. Note that the presence of unreacted MeOH can hamper the aminocarbonylation step resulting in an esterification (Scheme 6).75,76
Changing the substrate to 4-iodotoluene yielded 11f–11i, with a reduced yield for 11i. Using 4-iodobenzotrifluoride increased yields for 11j–11m with fully converted aniline and 4-fluoroaniline to 11j, 11l and 11m, with decreased esterification and maintaining good yields for 11n–11o. Note that 11l–11o were not described before. In contrast, 4-fluoroaniline and 4-fluoroiodobenzene gave significantly lower yields (11p). The use of 1-iodonaphthalene77 as the aryl halide afforded high yields (11q–11t), with excellent selectivity for naphthamides formation. We also exploited bromobenzene and chlorobenzene as aryl halide substrates. With aniline as substrate only traces of 11a were produced. However, 1-bromonaphthalene yielded 55% of 11p. Note that, in contrast, two literature reported examples for nitroarene methylation and a subsequent carbonylation using a palladium acetate/phosphine system under higher CO pressure proceeded via a two-step process.95
Supplementary information: details of the experimental procedures, characterization of the complexes and the DFT calculated details. See DOI: https://doi.org/10.1039/d5sc03892h.
| This journal is © The Royal Society of Chemistry 2025 |