Open Access Article
Anna
Lenarda
*a,
Itisha
Jain
a,
Aleksi
Kaleva
a,
Valtteri
Oksanen
b,
Sami
Heikkinen
a,
Risto
Koivula
a,
Tom
Wirtanen
b,
Michele
Melchionna
c,
Tao
Hu
d and
Juho
Helaja
*a
aDepartment of Chemistry, University of Helsinki, A. I. Virtasen aukio 1, P.O. Box 55, Finland. E-mail: anna.lenarda@helsinki.fi; juho.helaja@helsinki.fi
bVTT Technical Research Centre of Finland Ltd, P.O. Box 1000, FI-02044 Espoo, Finland
cDepartment of Chemical and Pharmaceutical Sciences, INSTM, University of Trieste, Via L. Giorgieri 1, 34127 Trieste, Italy
dResearch Unit of Sustainable Chemistry, Faculty of Technology, University of Oulu, 90014 Oulu, Finland
First published on 1st November 2025
Activated carbon derived porous materials, effectively enriched with OH and C
O groups, were found to mediate, in a cascade manner, the condensation between anilines and 3-hexenones or β-tetralones, followed by their aromatization to diarylamines. The reaction proceeds via in situ formation of enamine intermediates which are subsequently oxidatively dehydrogenated in presence of a molecular oxidant under inert atmosphere. The functional groups on the carbon surface contributed actively to the catalysis: phenolic hydroxyl groups were found to promote the coupling of amines and ketones to imines and their tautomerization to enamines, while the C
O groups of the quinoidic moieties catalyze the dehydrogenative aromatization step. The carbon material's extensive porous structure turns out to be critical to preserve the reactive β,γ-unsaturated cyclohexanone derivatives and their enamine intermediates from undesirable coupling and condensation side-reactions. The carbocatalyst can be regenerated by molecular N-oxo quinoline, which acts as a more convenient and cleaner stoichiometric oxidant in comparison with standard aerobic conditions (oxygen-rich atmosphere). The developed methodology delivered up to 93% yields for many diarylamines, formerly accessible exclusively via Pd-mediated couplings. Computational DFT study of possible enamine reaction modes with quinone model compounds, combined with kinetic isotope effects (KIE) suggest that the aromatization reaction is triggered by hydride abstraction at the benzylic position of the enamine intermediate.
Green foundation1. Our work demonstrates that porous, functionalized activated carbon can replace precious metal catalysts in C–N bond formation, enabling diarylamine synthesis via a cascade process under mild conditions. This advances green chemistry by providing a sustainable, metal-free alternative to Pd-mediated couplings with high selectivity and yields.2. We achieved efficient diarylamine synthesis (up to 93% yield) using a regenerable carbocatalyst and benign stoichiometric oxidant (N-oxo quinoline). This eliminates the need for transition metals and halogenated starting materials, while achieving high atom economy through a one-pot cascade reaction. 3. The process could be further improved by employing renewable feedstocks for the carbon material and by regenerating or recycling the stoichiometric oxidant. |
In contrast to this trend, we lately proved that air-oxidized commercial activated carbon (oAC) can aromatize partially hydrogenated hetero- and carbocycles. The developed catalytic protocol provides an attractive, alternative to Suzuki–Miyaura couplings, allowing a cost-effective, metal-free access to biaryls20 and heterobiaryls21via direct aerobic ODH aromatization. In this study, we employed oAC to promote a cascade route to diarylamines, relying on condensation and ODH aromatization of partially unsaturated cyclic ketones in the presence of anilines (Fig. 1D), where different functionalities on the carbon surface play different roles in the consequent steps of the reaction. In presence of heteroaromatic N-oxides as milder molecular oxidants under non-aerobic atmosphere, oxygen sensitive β,γ-arylfused and β,γ-unsaturated cyclohexenones, such as β-tetralone, were successfully converted into a wide range of diarylamines. While the large surface area and extensive porosity characteristic of AC materials, largely exploited for their high adsorption power, have been traditionally considered, for the same reason, cause of deactivation and therefore detrimental in catalysis,19 we found that these features stabilize herein sensitive substrates and intermediates, suppressing side reactions and enhancing overall efficiency.
Switching to Ar atmosphere and softer molecular oxidants turned out to be a first turning point. In this regard, our attention was captivated by reports that explored carbon materials’ ability to reduce some electron deficient organic compounds, like chlorinated nitroaryls.22 We first tested 1,4-dichloro-2-nitrobenzene 4 as terminal oxidant. At its best, the catalytic system including 1.5 equiv. of this compound delivered 55% yield of the aromatized product N-phenylnaphthalen-2-amine in 24 h at 140 °C under inert atmosphere using anisole as solvent (Fig. 2).
![]() | ||
| Fig. 2 Screening of various molecular oxidants, reaction conditions: catalyst loading (defined as 1 equiv.) 224 mg mmol−1, 0.1 mmol 1a, 0.17 mmol 2a, 0.1 mmol molecular oxidant in 1 mL of anisole. | ||
Encouraged by this result, we investigated a series of other known molecular oxidants using the same reaction conditions. Heteroaromatic N-oxides have been lately utilized as oxygen source in organic reactions,23 as oxidants in metal catalysed organic transformations24 and recognized as oxidants in biological redox reaction.25 Our screening of variously substituted pyridine N-oxides 5–8 with different electronic properties delivered the aromatized product in similar fair yields between 55 and 65% except in the case of 4-nitropyridine 1-oxide which, similarly to nitrosobenzene 9, failed to deliver any observable product (Fig. 2). Lastly, we tested quinoline 1-oxide 10, a widely employed precursor in organic synthesis that has been recently exploited as a nucleophilic oxidant in gold catalysed oxygenative transformations of unsaturated C–C bonds.26 Employing 1 equiv. of 10, we were able to obtain N-phenylnaphthalen-2-amine 3a in a satisfying 85% yield.
Any further variation of reaction conditions resulted in lower activities (Table 1). Lowering the temperature to 50 °C and running the reaction in toluene delivered 35% of the fully aromatized N-phenylnaphthalen-2-amine product 3a, together with 38% of the enamine intermediate N-phenyl-3,4-dihydronaphthalen-2-amine 3a′ in 24 h reaction time, while increasing the temperature to 90 °C only moderately improved the activity (50%). α,α,α-Trifluorotoluene, another high boiling solvent with higher polarity, delivered only 22% 3a′ and 22% 3a at 110 °C.
| Entry | Variation from optimized conditions | Yielda of 3a′ | Yielda of 3a |
|---|---|---|---|
| a Yield is calculated from quantitative 1H NMR (500 MHz, DMSO-d6) using 1,3,5-trimethoxybenzene as an external reference. b Optimized reaction conditions: catalyst loading (defined as 1 equiv.) 224 mg mmol−1, 0.1 mmol 1a, 0.17 mmol 2a, 0.1 mmol molecular oxidant in 1 mL of anisole. | |||
| 1 | — | 0% | 85% |
| 2 | 50 °C in toluene | 38% | 35% |
| 3 | 90 °C in toluene | 26% | 50% |
| 4 | 110 °C in trifluorotoluene | 22% | 22% |
| 5 | No oACair | 6% | 15% |
| 6 | No oACair, no oxidant (10) | 68% | 0% |
| 7 | 0.5 equiv.(112 mg mmol−1)b oACair | 0% | 78% |
| 8 | 2 equiv. (448 mg mmol−1)b oACair | 0% | 76% |
| 9 | 4 equiv. (896 mg mmol−1)b oACair | 0% | 78% |
As quinoline N-oxide can act as an oxidant on its own,24 we performed the reaction in absence of oACair to ensure its catalytic role. The yield obtained in this case was very low (15%), confirming its importance in the developed system. Partially reduced heteroaromatic systems can disproportionate, as reported for 1,2-dihydroquinolines,27 and the barrier for this reaction has been calculated to be relatively low for N-protonated species.28 To exclude the thermal occurrence of this process from intermediate 3a′ we ran the reaction in absence of both the carbon catalyst and the molecular oxidant: after 24 h at 140 °C, only 68% of 3a′ was observed, with no trace of aromatized product 3a excluding this reaction pathway.
Next, we proceeded to vary the catalyst loading: increasing the loading to 2, and even 4 equivalents, or lowering it to 0.5 equivalents lowered the yield similarly, even if not very significantly, to respectively 76, 78 and 78%, suggesting that 1 equivalent was already the optimal loading. We attribute the lower yield obtained with increased carbon loading to partial, irreversible adsorption of substrates or intermediates on the carbon surface.
The scope of the reaction was then studied, revealing that para-substituted anilines, both with electron donating and electron withdrawing groups, are, in general, quite reactive, with small variations in isolated yields (Fig. 3A, entries 2b–2d). Switching to N-heterocyclic arylamines, however, highlighted a distinct pattern: 2-aminopyridine (2g) and pyrimidin-4-amine (2f) showed the highest activity, which decreased in the case of N-(naphthalen-2-yl)pyridin-3-amine (3h), and even more drastically with N-(naphthalen-2-yl)pyrimidin-2-amine (3i), suggesting a relevant effect of the electronic properties of the arylamine partner.
Methoxy group substitution in different positions on the β-tetralone aromatic ring also affected the reactivity: high yield was obtained with 7-methoxy derivative (3j), while 6-methoxy or 5-methoxy β-tetralones turned out to be much less reactive (3l and 3m). The additional methyl substituent in 1-methyl-7-methoxy β-tetralone lowered the yield even further (3k). 6-Fluoro substituted β-tetralone produced the highest yield (3r). Swapping the aniline with pyrimidin-4-amine (2f), which previously showed higher activity did not change the yield obtained with 5-methoxy β-tetralone (3p).
To mimic a common substructure in pharmaceutical compounds, 3o, 7,8-dihydroquinolin-6(5H)-one (1g) was prepared via a reported procedure (see Supporting Information) and successfully reacted with pyrimidin-4-amine. The related diarylamine compound with aniline (3s) was also successfully isolated, although in lower yield.
The combination of electron withdrawing and electron donating substituents respectively in the β-tetralone and aniline starting materials resulted in moderate yields (3n and 3q). A larger polycyclic heteroaromatic dibenzo[b,d]furan-3-amine delivered 71% of aromatized product 3t when coupled with unsubstituted β-tetralone.
Additionally, two cyclic aliphatic amines, morpholine and piperidine, were tested for widening the scope to a different class of N-naphthalene amines. As a result, 4-(naphthalen-2-yl)morpholine 3u was successfully isolated in high yield, while only 35% of 1-(naphthalen-2-yl)piperidine 3v was obtained. Finally, employing secondary aromatic amine N-methyl aniline delivered 63% yield of product 3w.
Substituted 4-phenyl cyclohex-3-enones can be straightforwardly synthesized from bromoaryls and 1,4-dioxaspiro[4.5]decan-8-one via Grignard reaction followed by subsequent tandem elimination/deprotection (see SI). Interestingly, these compounds could be also used as partially aromatized building blocks to access 4-phenyl anilines, under the optimized conditions. When 4-phenylcyclohex-3-en-1-one 11a was used as reagent, (Fig. 3B) the product N-phenyl-[1,1′-biphenyl]-4-amine 12a was obtained in 55% yield, and no other product or leftover starting materials was observed in the crude mixture. Lowering the temperature to 110 °C and 90 °C, as well as changing the solvent from anisole to toluene, and all other attempts to optimize the reaction conditions resulted in lower yields (see SI, Table S1).
The scope of the reaction was explored varying the substituent in 4′ position of 4-phenylcyclohex-3-en-1-one and the aniline partner. 4′-Methyl and 4′-fluoro substituted starting materials delivered similar yield to the unsubstituted one when coupled with aniline (12b and 12c), while both electron rich and electron poor substituents on the amine lowered the yield to 40% (12e and 12f). Any attempt to use electron rich 4′-methoxy substituted aryl substrates produced a complex mixture of products, likely due to an additional coupling step of aromatized ring leading to formation of a mixture non-isolable side product (see SI, Fig. S1). In analogy with the previous scope, cyclic aliphatic amine morpholine was also tested and delivered respectively 40% and 58% yield of biphenyl amine products 12g and 12h when coupled respectively with 4′-methyl and unsubstituted 4-phenylcyclohex-3-en-1-one.
O, CO2H and OH groups of oACs (see Materials and methods section 3.1), and running the reaction of 1a and 2a in the same conditions with the modified catalysts (Fig. 5a).29,30 The derivatization of carboxylic acid groups with bromoacetophenone did not cause any major change in yield, suggesting their lack of active participation in the catalytic process. On the contrary, blocking carbonyls with phenyl hydrazine lowered the yield to 50%. The result became even more dramatic when the reaction was run in absence of the molecular oxidant quinoline 1-oxide (10), where only 44% of product could be observed. Interestingly, the esterification of phenolic groups had a considerable effect on the reaction, lowering the product yield to 66%. It seems therefore plausible that phenolic groups are somehow involved in the reaction mechanism, either by being converted to active carbonyls via oxidation, or as hydrogen bond donors in their original form.
As additional evidence of the contribution of o-quinone moieties to the catalysis, we selectively blocked them by modifying the carbocatalyst with 4,5-difluorobenzene-1,2-diamine, following a procedure reported by Fukushima et al.31 The effective chemical alteration of the desired functional groups was verified via XPS analysis (Fig. 4). As summarized in Fig. 4, after the 4,5-difluorobenzene-1,2-diamine labelling, the C
O component decreased from 2.2% to 0.4%, while two new signals appear: a lower, pyridinic component in the N 1s region and, more importantly, the characteristic F 1s signal at 687.0 eV.
![]() | ||
| Fig. 4 XPS analysis of oACair before and after 4,5-difluorobenzene-1,2-diamine labelling: summary of O1s deconvolutions (a) and (c), N 1s (b) and (d) and F 1s (e) peaks. Atomic values in Table in %. | ||
The reaction run using the F-labelled material as catalyst only delivered 48% of the aromatized product, further confirming the crucial role of o-quinone moieties in promoting the reaction. The role of surface chemistry in this catalytic process was explored in more depth using other reference AC materials (see Materials and methods section 3.1.3 for their preparation) characterized by different surface chemistry (Fig. 5b).21 Both oACair(Δ), oACair treated at 450 °C in inert atmosphere to fully remove carboxylic acid groups, and oACHNO3, which oxidation process in concentrated HNO3 at 140° C makes strongly acidic, did not significantly affect the reaction yield under standard conditions, confirming the low relevance of acidic functionalities for this reaction. Surprisingly, ACdm, simple commercial activated carbon washed with diluted HCl to remove metal impurities with no additional oxidative treatment, exhibited remarkable activity (67%) despite its lower content of surface oxygen groups, even in absence of molecular oxidant quinoline 1-oxide. Similarly, product 3a was obtained with 77% with oACair in absence of any oxidant. These findings suggest that carbonyl groups are not the sole factor in determining the catalytic activity, which instead is controlled by a more complicated mechanism than that of other ODH reactions.21 This idea was in agreement with the relatively low activity observed when we explored reduced graphene oxide (rGO), a widely reported catalyst in ODH,21 which instead, in this case gave only 50% yield of the main product together with a complex mixture of side products. The relatively high O content in this material, together with its different structural features compared to AC, made us want to investigate the role of porosity and surface area, as it appears to be a critical parameter for the studied reaction. We selected pyrene as an inert polycyclic aromatic compound to compete in occupying the pore volume with the substrates and intermediates. First, we ran the reaction in the presence of different amounts of pyrene, to physically block the pores of the appropriate size. We observed that the yield of 3a dropped to 57% and 50% in the presence of one and two equivalents of pyrene, respectively, as compared to the 85% obtained in absence of the additive. Contrarily, addition of 1 or 2 equivalent of pyrene in the reaction mixture when PQ, molecular model for o-quinones, was used as catalyst, did not alter its activity. Incidentally, the yield and product mixture obtained in these conditions is rather similar both to the one obtained with GO, and in the tests with pyrene blocking the pores of oACair. This seems to imply that the large surface area and extensive porosity, characteristic of AC materials, play a crucial role in promoting the reaction, either by stabilizing the ketone through π–π coordination or favoring its coupling with the aniline partner by increasing their local concentration (Fig. 5b).
![]() | ||
Fig. 5 (A) Catalyst test with oACair showing effect of selective blocking of functional groups, C O (without 10) refers to the experiment run with C O blocked catalyst in absence of molecular oxidant (B) Atomic percentages from XPS O 1s spectra (bars) and yield in 3a after 24 h, optimized conditions from Table 1 (dashed line with circles), ACdm, oACHNO3, oACair, and oACair(Δ), represent HCl-washed, HNO3 oxidized, air oxidized, and thermally treated air oxidized active carbons, respectively; rGO represents commercial reduced graphene oxide. (C) Catalyst test with oACair blocking pores with different pyrene loadings, and (D) Recycling of catalyst over four cycles. | ||
To confirm the effect of pyrene in altering the material's pore structure, we stirred oACair with different amounts of the probe molecule in the same conditions as we perform the catalytic test and, after thorough washing and drying (see Materials and methods, section 3.1), analyzed the samples by means of N2 physisorption (Fig. 5c). The clear correlation between pyrene loading and decreased surface area confirm its effectiveness in altering the pore structure of the material: the original surface area of over 1000 m2 g−1 becomes respectively 729 m2 g−1 and 662 m2 g−1 in presence of one or two equivalents of blocking probe, in good correlation with the 30% and 40% decrease in yield observed in the related catalytic experiments. A similar trend is evident in the decreased pore volumes (Fig. 5c). The BET analysis of rGO supports this hypothesis: its structural analogy to the pyrene blocked samples (Fig. 5c), results in similar behaviour in catalysis.
Finally, oACair was analyzed with solid-state 13C NMR (SI, Fig. S8). The spectrum is dominated by the strong band characteristic of graphitic C, which prevents detection of other functional groups.
We then assessed the robustness of the catalyst by carrying out recyclability tests. When oACair was used, after four sequential 6 h-long catalytic cycles, we did not observe any major decrease in activity, both for the synthesis of N-phenylnaphthalen-2-amine and N-phenyl-[1,1′-biphenyl]-4-amine (Fig. 5d).
Moreover, scaling up the reaction to 1 mmol scale resulted in no significant decrease of activity (isolated yield 75%). This aspect is encouraging from the point of view of industrial development. Inductively coupled plasma mass spectrometry (ICP-MS) was used to investigate the possible presence of residual metal impurities in the batch of activated carbon employed in this study. The analysis revealed a maximum Fe content of 600 ppm and notably lower levels of other metal impurities. Such trace concentrations are far below the threshold typically associated with catalytic activity, making trace-metal catalysis highly unlikely (reported in a previous publication, see ref. 21). Finally, the headspace of the reaction was analysed via MicroGC to evaluate the gaseous products composition. After 16 h reaction time no traces of H2 were detected, indicating against direct dehydrogenative regeneration of the carbocatalyst (see SI).
The rate of formation for the 3′ intermediates can be rationalized by comparing, on the one hand, the different amines’ basicity and nucleophilicity, that has been observed to be critical for imine formation rate,33 and on the other hand, the thermodynamic stabilities of enamines themselves. In fact, both these values increase going from 3g′ to 3a′ to 3u′ (Fig. 6a), in good correlation with what we observed experimentally.
The coupling and tautomerization steps are known to proceed swiftly in protic media under mildly acidic conditions,34 while without proton shift mediators these steps exhibit high barriers.35 However, the reaction proceeds smoothly delivering high yields in our optimized catalytic conditions, which make use of relatively dry non-protic solvents such as anisole and toluene.
Herein, we attribute this role to OH and CO2H functional groups abundantly present on activated carbon surfaces, which can mediate these steps even in the presence of minor amounts of moisture. Kinetic monitoring of the reaction with 1a and 2a using three different carbons as catalysts shows that, after 2 h, the enamine formation occurs the fastest with oACair, which carries an increased amount of these groups compared with ACdm, that can still mediate enamine formation rather well. The reaction, however, proceeds rather modestly with OH blocked oACair (Fig. 6b), underlining the relevant role of these groups in the formation of 3a′. To further explore this effect, we studied the reaction kinetics in presence of different, controlled amounts of water (Fig. 6c). Addition of 5 µl of H2O appears to have a beneficial effect, increasing the rate of the reaction significantly in the first hours. However, the product yield stops improving after 15 h, suggesting the occurrence of competing side reactions in these conditions. The effect is even more pronounced when the water addition is increased to 25 µl. Additionally, we carried out kinetic studies with selectively deuterated β-tetralones (Fig. 7). The only KIE (2.9) was received for 4-position deuterated ketone, which indicates that hydrogen/hydride abstraction at the benzylic position forms a reaction barrier affecting significantly the rate of the reaction. This agrees with computed rate-limiting barrier for the aromatization of enamine 3a′, that is H-4 abstraction by the quinone model (vide infra).
![]() | ||
| Fig. 7 Kinetic isotopic effect measurements; yield was determined from quantitative 1H NMR (500 MHz, DMSO-d6) of the crude product using 1,3,5-trimethoxybenzene as an external reference. | ||
First, we estimated the enamine formation reaction energy for 1a and amines 2a, 2g, 2u (see SI, Table S5) by applying explicit anisole molecules to dissolve H2O, obtaining exergonic energies for all conjugated enamines, −5.7, −7.9 and −3.1 (see SI) respectively for 3a′, 3g′ and 3u′, which are in good agreement with what we observed in the kinetic study.
Secondly, we became interested in the imine-amine tautomerization barrier: since the imine tautomer was not observed in the reaction monitoring, despite its formation preceding that of the enamine one, we supposed that a low activation energy for this process, in addition with its favorable thermodynamics, could kinetically explain this behavior. In absence of proton shift mediators, imine–enamine proton transfer scans returned remarkably high barriers (>60 kcal mol−1), significantly lowered by assisting molecules, as indicated by including two or three H2O or two H2O and one phenol (see SI). The calculations performed for imine 3a″ tautomerization to enamine 3a′ report, for example, lower activation energy barriers by assistance of two or three H2O or two H2O and one phenol, 28.8, 26.8 and 24.9 kcal mol−1, respectively (see SI, Fig. S9). A similar trend is observed for the exergonic thermodynamic stabilities of all enamines (see SI, Table S6): the barrier for the tautomerization of 3g″ to 3g′, for instance, if mediated with two H2O, is 31.3 kcal mol−1.
The activity of PQ as molecular catalyst for this reaction (see section 3.2) justifies its use also as an active site model for the quinoidic functionalities on the carbon surface. Quinones are capable to mediate both open and closed cell redox processes, though typically quinoidic oxidation of organic substrates proceeds via polar reaction mechanism involving hydride abstraction.36 KIE experiments of dihydropyridine compounds with strong hydride/electron donors by Fukuzumi et al. have shown that this process operates via ion-pair mechanism for electroneutral and mildly electron deficient p-benzoquinone derivatives, as demonstrated with chloranil and more electron rich quinones (E0(Q/Q˙−) < 0.01 V vs. SCE), while a radical mechanism is associated to strong electron acceptors, like 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ) (E0(Q/Q˙−) = 0.51 V vs. SCE).37 Meanwhile, by studying nicotinamide adenine dinucleotide models as electron/hydride donors and p-quinones as oxidants Cheng and coworkers discovered that empirical 1 V endothermic redox potential limits the radical pathway, while ionic hydride transfer mechanism can operate above this limit.38
The measured reduction potential for PQ, obtained via cyclic voltammetry, is −0.66 V vs. SCE (NBu4PF6 0.1 M in MeCN).39 Computation of oxidation potentials for the studied enamines results in values between 0.25–0.89 V, vs. SCE (MeCN) (Table S9, SI) indicating that the 1 V limit is exceeded for all the substrates apart for 3u′, which had, nonetheless, modest reactivity.
When we applied the computational concept that Floreancig and Liu recently used for benzylic ether C–H bond activation mechanism by DDQ40 to 3a′ and PQ we could confirm that single atom transfer (SET) and hydrogen atom transfer (HAT) are the less probable routes (see SI), while the C–H cleavage likely operates via hydride shift from the benzylic position of the enamine either in stepwise manner forming a carbocation intermediate (TS1, Fig. 8) or via concerted hydride – proton shift (TS2, Fig. 8).
The inspection of the potential energy surface calculated for the hydride shift from 3a′ to PQ (Fig. 8) shows that the initial van der Waals complex 3a′-PQ-vdW between 3a′ and PQ has an endergonic energy (ΔG = 1.8 kcal mol−1) exhibiting very weak charge transfer character (e3a′ = 0.02 and ePQ = −0.02). The reactivity from this point proceeds in polar fashion, and as the benzylic (position 4) carbocation of 3a′Fig. 6 Enamine 3a′ and PQ hydride shift energy profiles for 3a (kcal mol−1).is ca. 9 kcal mol−1 more stable than the allylic (position 3), we assume the PQ carbonyl O-attack takes place at this position. For the stepwise hydride shift pathway, in which the H-shift to oxygen is associated with secondary orbital interaction between the carbonyl carbon of PQ and the phenyl ring of 3a′ (see SI) we calculated 34.1 kcal mol−1 energy barrier TS1 and 25.3 kcal mol−1 for the endergonic reaction energy of intermediate 3a′BnC-HPQ-vdW, which, after a low barrier TS3 proton transfer step delivers the aromatized product 3a and H2PQ.
Alternatively, the reaction may proceed in a concerted manner, where the PQ O-attack takes place also at the benzylic position, and another H-shift to its carbonyl carbon is taking place in the same vibrational mode (TS2, Fig. 8). This route has a slightly higher energy barrier, TS2, (ΔG‡ = 35.8 kcal mol−1) and would deliver directly the aromatized product 3a together with a different acyloin-type isomer of hydroquinone, H2PQ′, which is known from previous studies to be smoothly oxidized back to quinone in presence of O2.41 Variation of the amine partner does not have a dramatic effect on this value, except in the case of 3g′ where the concerted pathway is the energetically preferred one (Table 2, entry 3).
| Entry | Compound (+H-bond donor) | TS1 | TS2 |
|---|---|---|---|
| The lower of the TS1/TS2 barriers is shown in bold, indicating the preferred stepwise or concerted route, respectively. | |||
| 1 | 3a′ + H2O | 30.8 | 34.2 |
| 2 | 3a′ + PhOH | 28.4 | 32.2 |
| 3 | 3g′ | 35.6 | 33.4 |
| 4 | 3u′ | 32.2 | 34.9 |
| 5 | 3j′ | 33.5 | 35.2 |
| 6 | 3l′ | 32.1 | 34.4 |
| 7 | 3m′ | 31.1 | 35.2 |
| 8 | 3a′/CQ instead of PQ | 31.8 | 32.9 |
Notably, when a water or phenol molecule H-bonds the oxygen of PQ, the reaction becomes less endergonic (Fig. S9, SI), with reaction barriers lowered to 30.8 and 28.4 kcal mol−1, respectively (Table 2), and intermediate charge transfer complexes (3a′BnC-HPQ-vdW) becoming more stable although still endergonic (21.2 and 14.4 kcal mol−1, respectively, see SI).
As PQ is a relatively small molecular quinone, we became interested in computing similar C–H activation routes for an o-quinone fused in larger polycyclic aromatic system that is structurally closer to graphene-like carbon materials. Analogous computations performed with coronene-1,2-dione (CQ) as a quinone model compound resulted in a similar energy profile as the one obtained for PQ (Fig. 9). Remarkably, the energy barriers were slightly lower for both the routes, being 31.8 and 31.9 kcal mol−1 for TS1 and TS2, respectively; while also the gap between them decreased to 1.1 kcal mol−1. Overall, with CQ as hydride abstractor, the bond forming and breaking distances were similar to those of PQ, however, the charge transfer character of complexes is slightly increased starting from initial vdW-complex (3a′-CQ-vdW, e3a′ = 0.09 and eCQ = −0.09). As the larger coronene platform offers a wider surface for vdW-interactions (NCI plot, SI) and extended conjugation for charge stabilization, the formation energy of the initial vdW-complex becomes exergonic (−1.4 kcal mol−1) and the intermediate complex, 3a′BnC-HCQ-vdW, 6 kcal mol−1 less endergonic (19.0 kcal mol−1). We presume that these factors operate similarly in the studied carbon materials carrying o-quinone edges, making both the hydride abstraction routes experimentally possible.
To conclude, we propose that the reaction could proceed on the carbocatalyst surface in a cascade manner (Fig. 10), initiated by coupling of a ketone with an amine, followed by imine condensation and imine–enamine tautomerization steps promoted by protic mediators. The aromatization is then triggered by quinoidic hydride abstraction and completed with deprotonation by hydroquinoidic anion moiety.
While the surface functionalities are key for triggering the catalytic transformation, the material's extensive porosity has the role of preserving sensitive starting material β-tetralone and stabilizing the enamine intermediates.
oAC HNO3 : ACdm (4.00 g) was placed in a flask and 8 mL of nitric acid (68% aq.) was slowly added forming a slurry. The flask was attached to a Dreschel bottle with NaOH aq., and then the flask was heated to 140 °C for 15 h. oAC was then dried in vacuum at 140 °C for 2 h, resulting in a yield of oACHNO3 of 3.88 g.
oAC air : ACdm (4.00 g) was placed in a porcelain crucible which was heated to 425 °C for 16 h (30 °C min−1) in an oven in presence of air. oACair was obtained in a yield of 3.11 g.
oAC air(Δ): oACair (3.11 g) was placed in a tubular oven under Ar flow (40 cm3 min−1) for 23 h. The temperature was increased from RT to 450 °C (10 °C min−1) and kept at 450 °C for 24 h while maintaining the Ar flow. After cooling to RT, oACair(Δ) was obtained in a yield of 3.05 g.
Phenolic hydroxyl groups were esterified with benzoic anhydride. Benzoic anhydride (5.0 g, 22.1 mmol) was dissolved to deoxygenated chloroform (50 mL), after which oACair (1.0 g) was added. Mixture was stirred at 60 °C for 24 hours. Modified carbocatalyst was filtered and washed with chloroform (2 L). Product was dried at 60 °C under vacuum for 24 hours.
Carboxylic acids were esterified with 2-bromo-1-phenylethanone. oACair (1.0 g) and 2-bromo-1-phenylethanone (2.0 g, 10 mmol) were mixed in deoxygenated chloroform (50 mL). Mixture was stirred for 5 h protected from light. Modified carbocatalyst was filtered and washed with chloroform (2 L). Product was dried at 60 °C under vacuum for 24 h.
F-labeling of o-quinone groups was carried out following a modified reported procedure.31 400 mg of oACair was dispersed in 40 ml of ethanol containing 144 mg (1 mmol) of 4,5-difluorobenzene-1,2-diamine. The reaction vessel was held at 60 °C overnight under an Ar atmosphere. Upon cooling the reaction mixture, the carbon powder was filtered and washed with 500 ml of ethanol and 200 ml of water. The carbon powder was subsequently dispersed in an aqueous 0.1 M HClO4 solution and stirred for 24 h. To avoid possible photoreactions, all reactions were conducted in the dark. Following acid treatment, the carbon powder was filtered and washed with 1000 ml of pure water and dried at 60 °C under vacuum for 24 h.
The catalytic behaviour and robustness of the oACs was confirmed by recyclability tests, which showed remarkable stability over four cycles. While carbocatalyst development studies typically focus on determining the active sites and increasing their abundance, here we identified and demonstrated the crucial role of the extended porous structure in the catalyst performance. Their stabilizing effect on both sensitive starting materials and intermediates prevents the occurrence of undesired side reactions, improving its selectivity and the overall efficiency of the process, as was demonstrated by pore blocking experiments with molecular probes.
We anticipate that this work will advance the development of carbon-based catalysts for ODH reactions, providing cost-effective and robust alternatives for transition metal mediated reactions. In particular, the established essential role of their pore structure for selectivity could be exploited to explore new types of reactivity in these materials.
| This journal is © The Royal Society of Chemistry 2026 |