Skip to main content
ACS Omega logoLink to ACS Omega
. 2020 Mar 27;5(13):7576–7583. doi: 10.1021/acsomega.0c00402

Highly Selective Synthesis of Hydrazoarenes from Nitroarenes via Polystyrene-Supported Au-Nanoparticle-Catalyzed Reduction: Application to Azoarenes, Aminoarenes, and 4,4′-Diaminobiaryls

Jee Eun Hong 1, Yeonghun Jung 1, Youmie Park 1, Yohan Park 1,*
PMCID: PMC7144144  PMID: 32280901

Abstract

graphic file with name ao0c00402_0001.jpg

A selective synthesis of hydrazoarene from nitroarene and its application are reported. Using polystyrene (PS) resins as solid supports for Au nanoparticles (AuNPs), polystyrene-supported Au nanoparticles (AuNPs@PS) were synthesized and characterized. In the presence of AuNPs@PS (1.0 mol %) as a catalyst, nitroarenes afforded corresponding hydrazoarenes (up to 99%) with high selectivity (up to 100%) under mild reaction conditions (NaBH4, 50% aq. EtOH, and room temperature). Depending on the reaction conditions (the amount of NaBH4, the substituent of nitroarenes, and the sequential addition of HCl), nitroarenes were converted to corresponding azoarenes (up to 95%), aminoarenes (up to 99%), and 4,4′-diaminobiaryls (up to 99%). Our easily recyclable catalytic system using a solid-phase reaction vessel provides an attractive synthetic method in an eco-friendly and sustainable manner.

Introduction

Au nanoparticles (AuNPs) have received continuous attention as nanosized material-based catalysts because they confer the advantages of both homogeneous and heterogeneous catalysis.1,2 AuNPs have a very high surface-to-volume ratio to enhance effective collisions between reactants and catalysts. At the same time, AuNPs are easily separated from the reaction mixture because they are insoluble.3 The catalytic activity of Au particles is directly related to their nanometer length scale.4 Considering that Au has long been known as an inert metal, this is a very interesting feature.

To minimize size growth of AuNPs,5 metal oxides,69 graphene derivatives,10,11 porous silicas,12 and polymers13,14 have been commonly used as solid supports. These supports can increase the catalytic activity through the synergistic effect between supports and AuNPs,15 as well as the chemical stability in reaction solutions by inhibiting the aggregation and precipitation of AuNPs.4 Polystyrene (PS) offers the advantages of supporting AuNPs of various sizes, functional group diversity, and chemical stability and is easy to handle and recover through simple filtration.16 In particular, the use of solid-phase reaction vessels17,48 with polystyrene-supported AuNPs (AuNPs@PS) can contribute to eco-friendly and sustainable chemistry by maximizing catalyst recovery and reuse. Reported examples using AuNPs@PS as a catalyst in organic reactions include lactonization of diol,18 oxidation of 1,4-dioxane to 1,4-dioxan-2-ol,19 reduction of nitro to amine,20 Suzuki–Miyaura coupling,20 and hydration of nitrile to amide.21

The reduction of nitroarenes is an interesting reaction because it proceeds through a more complicated route than expected and can produce a variety of intermediates, such as hydrazoarenes, azoarenes, and aminoarenes. These are essential intermediates in the manufacture of pharmaceuticals and dyes.22,23 According to the mechanism proposed by Haber,2427 nitroarenes are reduced to aminoarenes via the direct route through nitrosoarenes and N-arylhydroxylamines or the condensation route through azoxyarenes, azoarenes, and hydrazoarenes. Since these various species can be present in the reaction mixture at the same time, selective reduction toward a specific product is of paramount importance for the success of this reaction. Therefore, several cases have been reported in which newly developed nanosized material-based catalysts were applied to the selective reduction of nitroarenes to hydrazoarenes,28 azoarenes,15,29 azoxyarenes,30 aminoarenes,3032 and N-arylhydroxylamines.30,33

Among them, hydrazoarenes are useful intermediates that can be converted to azoarenes, aminoarenes, 4,4′-diaminobiaryls, and 2-aminodiarylamines in one simple step.34 Particularly, 4,4′-diaminobiaryls generated from hydrazoarenes via 5,5-sigmatropic rearrangement contain the biphenyl moiety, one of the most important building blocks in chemistry.35 However, relatively few studies have been reported that address the synthesis of hydrazoarenes from nitroarenes as the main products compared to examples of azoarenes and aminoarenes. In this paper, we reported the selective synthesis of hydrazoarenes from corresponding nitroarenes via catalytic reduction by AuNPs@PS (1.0 mol %). Reactions were carried out under mild reducing conditions (NaBH4, 50% aq. EtOH, room temperature) in a solid-phase reaction vessel.

Results and Discussion

To synthesize AuNPs@PS, anion-exchange PS resins were selected since we expected that cations (RNMe3+) of resins interact with anions (BH4 and AuCl4) to help AuNPs form at the PS resin. Among the resins, PS1 was chosen because it has been used as a solid support for metal NPs in several studies.21,33,3639 Also, we thought that a smaller particle size and lower cross-linking of PS can provide a higher catalytic performance to the AuNPs@PS hybrid based on the previous report.40 Therefore, we chose PS2 and PS3, and detailed properties of PS1–3 obtained from commercial sources are shown in Table S1.48

Following the previous protocol for polystyrene-supported metal NPs,21,33,3639 AuNPs@PS1–3 were synthesized in two steps using PS1–3 (Scheme 1). The first step was the anion exchange of chloride to borohydride, and the second step was to reduce chloroauric acid to AuNPs in DMF at 100 °C. DMF is a polar aprotic solvent. Polar solvents can induce more swelling of PS1–3 with quaternary ammonium functional groups than nonpolar solvents.41 An aprotic solvent was preferred to avoid an acid–base reaction caused by a protic solvent. The reaction temperature (100 °C) was determined by comparing rate constants of synthesized AuNPs@PS hybrids through the initial screening by a pseudo-first-order kinetic model used in our previous work11 (data not shown).

Scheme 1. Preparation of Polystyrene-Supported Au Nanoparticles.

Scheme 1

Synthesized AuNPs@PS1–3 were examined by scanning electron microscopy (SEM), and the overall sizes of AuNPs@PS1–3 were 831, 122, and 69 μm, respectively (Figure 1). High-magnification SEM images of the AuNPs@PS1–3 surfaces are presented in Figure S2.48 Additionally, the hybrids were ground with a mortar and pestle and measured by field-emission transmission electron microscopy (FE-TEM) to determine the sizes of NPs. Average particle sizes of Au in corresponding AuNPs@PS1–3 were 9.4, 17.6, and 6.8 nm (Figure 2). Among the three hybrids, AuNPs@PS3 exhibited the smallest overall size and average AuNP size. Prior to subsequent experiments, the loading of Au on every hybrid was determined by inductively coupled plasma mass spectrometry (ICP-MS), and it was confirmed that the Au loading on AuNPs@PS1–3 is 0.05 mmol/g. This indicates that Au was quantitatively loaded on PS1–3 during the second step in Scheme 1.

Figure 1.

Figure 1

SEM images of (a) AuNPs@PS1 (∼831 μm), (b) AuNPs@PS2 (122 ± 28 μm), and (c) AuNPs@PS3 (69 ± 16 μm).

Figure 2.

Figure 2

FE-TEM images and Au nanoparticle size distributions of (a) AuNPs@PS1 (9.4 ± 3.2 nm), (b) AuNPs@PS2 (17.6 ± 5.0 nm), and (c) AuNPs@PS3 (6.8 ± 2.2 nm).

The AuNPs@PS1–3 hybrids of the same weight (50 mg) were tested to study the reaction conditions for hydrazobenzene (2a, Table 1, entries 1–3). The substrate was nitrobenzene (1a, 0.2 M), and NaBH4 (10 equiv.) was used as a reducing agent in 50% aq. EtOH at room temperature. The amount of Au on 50 mg of AuNPs@PS1–3 is 1.0 mol % substrate 1a (0.25 mmol). Conversion yields were calculated by 1H NMR based on the weight of crude products and the molar ratio of each included species.42,48 All three catalysts produced hydrazobenzene (2a) with very high yields (entry 1, AuNPs@PS1, 97%; entry 2, AuNPs@PS2, 99%; entry 3, AuNPs@PS3, 99%) and selectivity (100%), but AuNPs@PS3 exhibited the shortest reaction time (4 h) and was selected as the optimal catalyst (entry 3). The highest catalytic performance of AuNPs@PS3 is thought to be due to the smallest overall size and average AuNP size and the lowest cross-linking of PS among three hybrids. Interestingly, AuNPs@PS2 with the largest average AuNP size showed a slight increase in the yield (entry 2, 99%) compared to AuNPs@PS1 (entry 1, 97%). Although these results are very limited, they support that a smaller particle size and lower cross-linking of PS can give a higher catalytic performance to the AuNPs@PS hybrid. Using half as much NaBH4 (5 equiv.) as that in entry 3 still afforded hydrazobenzene (2a) with high yield (99%) and selectivity (100%) in entry 4. Even with 25 mg of AuNPs@PS3 (entry 5) and the increased reaction time from 6 to 10 h, the yield (98%) and selectivity (100%) were almost maintained compared to entry 4. Interestingly, 2.5 equiv. of NaBH4 gave azobenzene (3a) with a 76% yield and 97% selectivity (entry 6). The results for entries 4 and 6 reveal that hydrazobenzene (2a) or azobenzene (3a) can be selectively synthesized by controlling the amount of NaBH4 as a reducing agent. When only a single solvent was used—EtOH or H2O—different main products were obtained depending on the solvent (entry 7, EtOH, 4a; entry 8, H2O, 5a), but the yield (entry 7, 59%; entry 8, 55%) and selectivity (entry 7, 63%; entry 8, 70%) were relatively low compared to 50% aq. EtOH. The results for entries 7 and 8 indicate that the composition of the solvent is a very important factor in controlling the selectivity of the reaction. Additionally, it was confirmed that no reduction reaction of nitrobenzene (1a) occurred without AuNPs@PS3 (entry 9) or with PS3-BH4 alone (entry 10). This shows that AuNPs are a key site among the AuNPs@PS3 hybrids, which exhibits the catalytic activity. When 1.0 mol % HAuCl4 was used, azoxybenzene (4a) was obtained as the main product (entry 11). Since the reaction was a reducing system, perhaps, colloidal AuNPs were generated, and the reaction from nitrobenzene (1a) to azoxybenzene (4a) proceeded.

Table 1. Study on the Reaction Conditions for Reduction of Nitrobenzene (1a) under NaBH4 to Hydrazobenzene (2a), Azobenzene (3a), Azoxybenzene (4a), and Aniline (5a).

graphic file with name ao0c00402_0006.jpg

          yield (%)b
 
entry catalysta NaBH4 (equiv.) solvent time (h) 2a 3a 4a 5a selectivity (%)c
1 AuNPs@PS1 10 50% aq. EtOH 24 97       100
2 AuNPs@PS2 10 50% aq. EtOH 24 99       100
3 AuNPs@PS3 10 50% aq. EtOH 4 99       100
4 AuNPs@PS3 5.0 50% aq. EtOH 6 99       100
5d AuNPs@PS3 5.0 50% aq. EtOH 10 98       100
6 AuNPs@PS3 2.5 50% aq. EtOH 24 2 76     97
7 AuNPs@PS3 5.0 EtOH 24 5 29 59   63
8 AuNPs@PS3 5.0 H2O 24 20 4   55 70
9   5.0 50% aq. EtOH 24 nre  
10 PS3-BH4 5.0 50% aq. EtOH 24 nr  
11f HAuCl4 5.0 50% aq. EtOH 24 1 7 68   89
a

50 mg of catalysts (1.0 mol %) was used.

b

Yields were determined by 1H-NMR spectroscopy.

c

Selectivity was calculated based on the major product.

d

25 mg of AuNPs@PS3 (0.5 mol %) was used.

e

nr means no reaction.

f

1.0 mol % HAuCl4 was used.

The optimal catalyst, AuNPs@PS3, was examined by TEM energy-dispersive X-ray spectroscopy (EDS) for elemental analysis (Figure 3), and high-resolution X-ray diffraction (HR-XRD) (Figure 4), high-resolution TEM (HR-TEM) (Figure 5a), and selected area electron diffraction (SAED) (Figure 5b) were employed for the crystalline structure analysis. TEM EDS mapping images showed C, N, and Au atoms at a single point of the hybrid. Based on the results of XRD, HR-TEM, and SAED, the crystalline nature of AuNPs@PS3 exhibited a face-centered cubic structure.

Figure 3.

Figure 3

TEM-EDS mapping images of AuNPs@PS3: C (green), N (blue), and Au (red).

Figure 4.

Figure 4

XRD pattern of AuNPs@PS3.

Figure 5.

Figure 5

HR-TEM images: (a) d-spacing and (b) SAED pattern of AuNPs@PS3.

The mechanism that occurs on the surface of AuNPs as a catalyst is not yet known. The Langmuir–Hinshelwood model might be one way of explaining the bimolecular reaction on the surface of AuNPs.43,44 According to the description of this model, BH4 is adsorbed on the surface of AuNP in AuNPs@PS3 to form AuNP-H. After nitrobenzene (1a) is adsorbed here again, the reduction reaction proceeds by AuNP-H.28,45

Hydrazobenzene (2a), azobenzene (3a), azoxybenzene (4a), nitrosobenzene (6a), and N-phenylhydroxylamine (7a) were reacted under the conditions of (i) and (ii) to trace the reduction pathway of nitrobenzene (1a) (Scheme 2). The reaction conditions (i) and (ii) are the same as those of entries 4 and 8 of Table 1, respectively, and the difference between the two conditions is whether 50% aq. EtOH or H2O is used as a solvent. Hydrazobenzene (2a) did not afford aniline (5a), and hydrazobenzene (2a) was recovered quantitatively in both conditions (Scheme 2a,b). Azobenzene (3a) provided hydrazobenzene (2a, 1 h, 99%) in 50% aq. EtOH (Scheme 2c), but no reaction occurred without the catalyst, and 99% of azobenzene (3a) was recovered. Azoxybenzene (4a) and nitrosobenzene (6a) both provided hydrazobenzene (2a) in 50% aq. EtOH (Scheme 2d, 1.5 h, 98%; Scheme 2e, 3 h, 99%). Interestingly, N-phenylhydroxylamine (7a) directly afforded hydrazobenzene (2a, 3 h, 82%) without any side products in 50% aq. EtOH (Scheme 2f), while aniline (5a) did not change to any other compounds (data not shown). This is probably because N-phenylhydroxylamine (7a) is in equilibrium with nitrosobenzene (6a) to produce azoxybenzene (4a), and then azoxybenzene (4a) is reduced to hydrazobenzene (2a) through the condensation route. However, N-phenylhydroxylamine (7a) yielded hydrazobenzene (2a, 4%), azobenzene (3a, 69%), azoxybenzene (4a, 4%), and aniline (5a, 23%) in H2O (Scheme 2g). When only H2O was used as a solvent, the reaction proceeded through both direct and condensation routes, and aniline (5a) was produced from N-phenylhydroxylamine (7a) rather than hydrazobenzene (2a). Similar solvent-dependent changes between the direct and condensation routes of nitrobenzene reductions have been found in a recent paper.30

Scheme 2. Reductions of (a) Hydrazobenzene (2a) under (i), (b) Hydrazobenzene (2a) under (ii), (c) Azobenzene (3a) under (i), (d) Azoxybenzene (4a) under (i), (e) Nitrosobenzene (6a) under (i), (f) N-Phenylhydroxylamine (7a) under (i), and (g) N-Phenylhydroxylamine (7a) under (ii).

Scheme 2

Reaction conditions: (i) AuNPs@PS3 (1.0 mol %), NaBH4 (5 equiv.), 50% aq. EtOH, rt; (ii) AuNPs@PS3 (1.0 mol %), NaBH4 (5 equiv.), H2O, rt.

The proposed mechanism for synthesis of hydrazobenzene (2a) from nitrobenzene (1a) under the conditions of (i) is shown in Scheme 3. Reduction of N-phenylhydroxylamine (7a) to aniline (5a) seems to be slower than condensation of nitrosobenzene (6a) and N-phenylhydroxylamine (7a) to azoxybenzene (4a). The fact that nitrosobenzene (6a) and N-phenylhydroxylamine (7a) were not observed by 1H NMR indicates that condensation to azoxybenzene (4a) proceeds very quickly. The recently published literature also showed a similar phenomenon.28 Thus, the reaction proceeded through the condensation route, and the route was stopped at hydrazobenzene (2a) so that 2a could be obtained with a high yield (99%) and selectivity (100%).

Scheme 3. Proposed Mechanism.

Scheme 3

Various nitroarenes 1 with 1.0 mol % AuNPs@PS3 calculated based on the moles of Au were tested to determine the scope and limitations of the reaction (Table 2). Nitroarenes 1ag afforded 14 corresponding hydrazoarenes 2ag (up to 99%). In general, the synthesis of substituted hydrazobenzenes 2bg proceeded more slowly than hydrazobenzene (2a), so 10 equiv. of NaBH4 was used. A trend was observed that showed shorter reaction times and higher yields in the order of m- (2bB, 16 h, 98%; 2 dB, 4 h, 99%), o- (2bA, 24 h, 98%; 2dA, 24 h, 85%), and p-substituted hydrazoarenes (2bC, 24 h, 89%; 2dC, 24 h, 69%). A similar tendency was reported in the previous literature.15 It is thought that the nitro group and its intermediates in m-substituted nitroarenes are less affected by the electronic effect of substituents in the synthesis of hydrazoarenes. Among m-substituted nitroarenes, longer reaction times and lower yields were observed in the order of F, Cl, and Br (2dB, 4 h, 99%; 2eB, 6 h, 78%; 2fB, 24 h, 69%). The steric effect of substituents seems to be also important in this reaction.

Table 2. Synthesis of hydrazoarenes 2a.

graphic file with name ao0c00402_0005.jpg

graphic file with name ao0c00402_0004.jpg

a

Isolated yields.

b

5.0 equiv. of NaBH4 was used.

To investigate further scopes and limitations, azoarenes 3 were synthesized by adjusting the amount of NaBH4 (Table 3). Reactions were carried out using 5.0 equiv. of NaBH4 in the presence of AuNPs@PS3 (1.0 mol %). Among methylated azobenzenes 3bAC, m-methylated azobenzene 3bB showed the lowest yield (39%). Although the chemical yield of p-methylated azobenzene 3bC (18 h, 88%) was slightly lower than that of o-methylated azobenzene 3bA (24 h, 94%), less hydrazoarenes were observed on the TLC plate during the reaction, so additional p-substituted azoarenes 3dC, 3eC, and 3gC were synthesized. Seven azoarenes 3 were synthesized (up to 95%). Interestingly, it was found that o- (1iA), m- (1iB), and p-nitrophenol (1iC) gave high yields of corresponding aminoarenes (5iA, 99%; 5iB, 98%; 5iC, 99%) in a relatively short reaction time (2 h). This is probably because the reaction proceeded through the direct route due to the repulsive effect between corresponding nitrosophenolate and hydroxyaminophenolate groups, which can be generated from nitrophenols (1iA, pKa = 7.23; 1iB, pKa = 8.36; 1iC, pKa = 7.15)46 under basic conditions (pH = 11.2). The fact that p-nitrobenzoic acid (1jC, pKa = 3.44)47 afforded p-aminobenzoic acid (5jC, 6 h, 17%) supports this possibility because p-nitrobenzoic acid (1jC) can also be converted to carboxylate intermediates in the reaction media. These limited exceptions toward aminoarenes suggest that the type of substituent is an important factor in determining the direct or condensation route.

Table 3. Synthesis of Azoarenes 3 or Aminoarenes 5a.

graphic file with name ao0c00402_0003.jpg

graphic file with name ao0c00402_0002.jpg

a

Isolated yields.

b

2.5 equiv. of NaBH4 was used.

c

Conversion yield determined by 1H-NMR spectroscopy using 1,3,5-trimethoxybenzene as an internal standard.

The recyclability test of the AuNPs@PS3 catalyst was performed using nitrobenzene (1a) and a solid-phase reaction vessel. In a total of 10 cycles, hydrazobenzene (2a) was synthesized without a significant decrease in yields (99–90%) (Figure 6). After the recyclability test, AuNPs@PS3 was filtered, and any traces of Au in the filtrate were measured by ICP-MS: they were found to be below the detection limit.48 This proves that no Au leached from the PS3 solid support during the reaction and that the reaction followed a heterogeneous catalytic pathway. It also indicates that the loading of Au on the recovered catalyst was maintained. However, the average size of AuNPs on the recovered catalyst increased from 6.8 to 9.8 nm as shown in Figure S12.48 In the recyclability test, the decrease in yields might be due to the size growth of AuNPs. The same reaction strategy was also employed for gram-scale synthesis of hydrazobenzene (2a).48 It exhibited a slightly decreased chemical yield (84%) but can still be applied for industrial mass production.

Figure 6.

Figure 6

Recyclability of AuNPs@PS3.

The one-pot synthesis of 4,4′-diaminobiaryls 8 from corresponding nitroarenes 1 was carried out (Scheme 4). Nitroarenes 1 were reacted with AuNPs@PS3 for an appropriate time under the conditions of entry 4 in Table 1 followed by addition of concentrated hydrochloric acid. After 4 h, corresponding 4,4′-diaminobiaryls 8 were successfully synthesized (up to 99%).

Scheme 4. One-Pot Synthesis of 4,4′-Diaminobiaryls 8 from Nitroarenes 1.

Scheme 4

Conclusions

In conclusion, a novel synthetic method for hydrazoarenes (with up to 99% yield and 100% selectivity) from nitroarenes using polystyrene-supported Au nanoparticles as catalysts (1.0 mol %) was developed. It is notable that the amount of NaBH4, the type of solvent, and the substituent of nitroarenes play integral roles for selective reduction toward a specific product. As an application of the developed method, nitroarenes afforded azoarenes (up to 95%), aminoarenes (up to 99%), and 4,4′-diaminobiaryls (up to 99%) depending on the amount of NaBH4, the substituents of nitroarenes, and the sequential addition of HCl, respectively. Our new catalytic system using recyclable catalysts, mild reducing conditions, green solvents, and solid-phase reaction vessels provides eco-friendly, sustainable, and industrially favorable synthetic methods for hydrazoarenes, azoarenes, aminoarenes, and 4,4′-diaminobiaryls. The synthetic potential of this method for a continuous flow chemistry system is now under investigation.

Experimental Section

Preparation of PS-BH4

In a round-bottom flask (100 mL), distilled water (30 mL) was added to sodium borohydride (100 mg) and PS1–3-Cl (4.0 g). After stirring for 5 h at room temperature, the reaction mixture was filtered under reduced pressure, washed with distilled water until pH ∼7, and then washed with acetone. The obtained resin (PS1–3-BH4) was dried in vacuo for at least 6 h.

Preparation of AuNPs@PS1–3

In a round-bottom flask (50 mL), a stock solution of HAuCl4·3H2O in DMF (50 mM, 1.5 mL) was added to PS1–3-BH4 (1.5 g) and DMF (13.5 mL). The mixture was vigorously stirred for 1 h at 100 °C until the pale-yellow surface of PS1–3 turned completely reddish purple. Thereafter, the resin beads were filtered under reduced pressure, washed with distilled water, and then washed with acetone. The obtained AuNPs@PS1–3 were dried in vacuo for at least 6 h.

Calculation of Au Loading on AuNPs@PS1–3

To determine the initial concentration of Au in ppb, a stock solution of HAuCl4·3H2O in DMF (50 mM, 1.5 mL) was added to DMF (13.5 mL) in a 20 mL vial and measured by ICP-MS. After completion of the AuNPs@PS1–3 preparation, the drained filtrate in the 20 mL vial was also measured by ICP-MS to confirm the concentration of unreacted Au in ppb. The Au loading on AuNPs@PS1–3 was calculated based on the equation below {nAu: mole of added HAuCl4·3H2O (mmol), mAuNPs@PS: weight of synthesized AuNPs@PS1–3 (g), c1: initial concentration of Au (ppb), c2: concentration of unreacted Au (ppb)}:

graphic file with name ao0c00402_m001.jpg 1

General Procedure for Hydrazoarenes 2

In a solid-phase reaction vessel (5.0 g), sodium borohydride (47.3 mg, 1.25 mmol), AuNPs@PS3 (50 mg, 1.0 mol %), and 50% aq. ethanol (1.25 mL) were added together. Then nitrobenzene (1a, 25.7 μL, 0.25 mmol) was added to the reaction mixture. The reaction was monitored by TLC. After stirring for 6 h at room temperature, the filtrate was drained using water and dichloromethane for recovery of AuNPs@PS3. The filtrate was extracted with dichloromethane (10 × 3 mL). The organic layer was washed with brine (3 × 3 mL), dried over anhydrous MgSO4, filtered, and concentrated in vacuo. The residue was purified on silica gel (petroleum ether:DCM = 10:1 → 1:1) to afford hydrazobenzene (2a) as a pale-yellow solid (22.8 mg, 99% yield).

Acknowledgments

This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Republic of Korea (NRF-2016R1D1A1B04930774).

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.0c00402.

  • Chemicals, analytical equipment, reaction vessels, experimental details, and NMR spectra for synthesized compounds (PDF)

The authors declare no competing financial interest.

Supplementary Material

ao0c00402_si_001.pdf (3.3MB, pdf)

References

  1. Corma A. Attempts to fill the gap between enzymatic, homogeneous, and heterogeneous catalysis. Catal. Rev. 2011, 46, 369–417. 10.1081/CR-200036732. [DOI] [Google Scholar]
  2. Shi F.; Tse M. K.; Pohl M.-M.; Brückner A.; Zhang S.; Beller M. Tuning catalytic activity between homogeneous and heterogeneous catalysis: improved activity and selectivity of free nano-Fe2O3 in selective oxidations. Angew. Chem., Int. Ed. 2007, 46, 8866–8868. 10.1002/anie.200703418. [DOI] [PubMed] [Google Scholar]
  3. Polshettiwar V.; Varma R. S. Green chemistry by nano-catalysis. Green Chem. 2010, 12, 743–754. 10.1039/b921171c. [DOI] [Google Scholar]
  4. Corma A.; Garcia H. Supported gold nanoparticles as catalysts for organic reactions. Chem. Soc. Rev. 2008, 37, 2096–2126. 10.1039/b707314n. [DOI] [PubMed] [Google Scholar]
  5. Stratakis M.; Garcia H. Catalysis by supported gold nanoparticles: beyond aerobic oxidative processes. Chem. Rev. 2012, 112, 4469–4506. 10.1021/cr3000785. [DOI] [PubMed] [Google Scholar]
  6. Raptis C.; Garcia H.; Stratakis M. Selective Isomerization of Epoxides to Allylic Alcohols Catalyzed by TiO2-Supported Gold Nanoparticles. Angew. Chem., Int. Ed. 2009, 48, 3133–3136. 10.1002/anie.200805838. [DOI] [PubMed] [Google Scholar]
  7. Grirrane A.; Corma A.; García H. Gold-catalyzed synthesis of aromatic azo compounds from anilines and nitroaromatics. Science 2008, 322, 1661–1664. 10.1126/science.1166401. [DOI] [PubMed] [Google Scholar]
  8. Zhou N.; Polavarapu L.; Wang Q.; Xu Q.-H. Mesoporous SnO2-coated metal nanoparticles with enhanced catalytic efficiency. ACS Appl. Mater. Interfaces 2015, 7, 4844–4850. 10.1021/am508803c. [DOI] [PubMed] [Google Scholar]
  9. Chaiseeda K.; Nishimura S.; Ebitani K. Gold Nanoparticles Supported on Alumina as a Catalyst for Surface Plasmon-Enhanced Selective Reductions of Nitrobenzene. ACS Omega 2017, 2, 7066–7070. 10.1021/acsomega.7b01248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Pocklanova R.; Rathi A. K.; Gawande M. B.; Datta K. K. R.; Ranc V.; Cepe K.; Petr M.; Varma R. S.; Kvitek L.; Zboril R. Gold nanoparticle-decorated graphene oxide: synthesis and application in oxidation reactions under benign conditions. J. Mol. Catal. A: Chem. 2016, 424, 121–127. 10.1016/j.molcata.2016.07.047. [DOI] [Google Scholar]
  11. Choi D. H.; Hong J. E.; Jung Y.; Park H. I.; Baek S. T.; Kim S. J.; Lee C. G.; Park Y.; Park S. J.; Park Y. Synthesis of gold nanoparticles supported at graphene derivatives using green reductants and evaluation of their catalytic activity in 4-nitrophenol reduction. Gold Bull. 2019, 52, 165–174. 10.1007/s13404-019-00258-w. [DOI] [Google Scholar]
  12. Qureshi Z. S.; Sarawade P. B.; Hussain I.; Zhu H.; Al-Johani H.; Anjum D. H.; Hedhili M. N.; Maity N.; D’Elia V.; Basset J. M. Gold Nanoparticles Supported on Fibrous Silica Nanospheres (KCC-1) as Efficient Heterogeneous Catalysts for CO Oxidation. ChemCatChem 2016, 8, 1671–1678. 10.1002/cctc.201600106. [DOI] [Google Scholar]
  13. Wang S.; Zhang J.; Yuan P.; Sun Q.; Jia Y.; Yan W.; Chen Z.; Xu Q. Au nanoparticle decorated N-containing polymer spheres: additive-free synthesis and remarkable catalytic behavior for reduction of 4-nitrophenol. J. Mater. Sci. 2015, 50, 1323–1332. 10.1007/s10853-014-8692-3. [DOI] [Google Scholar]
  14. Sasidharan M.; Anandhakumar S.; Bhanja P.; Bhaumik A. Highly efficient Au hollow nanosphere catalyzed chemo-selective oxidation of alcohols. J. Mol. Catal. A: Chem. 2016, 411, 87–94. 10.1016/j.molcata.2015.10.007. [DOI] [Google Scholar]
  15. Liu X.; Li H. Q.; Ye S.; Liu Y. M.; He H. Y.; Cao Y. Gold-catalyzed direct hydrogenative coupling of nitroarenes to synthesize aromatic azo compounds. Angew. Chem., Int. Ed. 2014, 53, 7624–7628. 10.1002/anie.201404543. [DOI] [PubMed] [Google Scholar]
  16. Liguori F.; Moreno-Marrodan C.; Barbaro P. Metal nanoparticles immobilized on ion-exchange resins: A versatile and effective catalyst platform for sustainable chemistry. Chin. J. Catal. 2015, 36, 1157–1169. 10.1016/S1872-2067(15)60865-8. [DOI] [Google Scholar]
  17. Merrifield R. B. Solid phase peptide synthesis. I. The synthesis of a tetrapeptide. J. Am. Chem. Soc. 1963, 85, 2149–2154. 10.1021/ja00897a025. [DOI] [Google Scholar]
  18. Oberhauser W.; Lavacchi A.; Vizza F.; Capozzoli L.; Lee H. M. Aerobic diol lactonization by Au-nanoparticles supported onto an anion-exchange resin. Appl. Catal. A–Gen. 2013, 451, 58–64. 10.1016/j.apcata.2012.11.012. [DOI] [Google Scholar]
  19. Sharma A. S.; Kaur H. Au NPs@polystyrene resin for mild and selective aerobic oxidation of 1,4 dioxane to 1,4 dioxan-2-ol. Catal. Commun. 2017, 90, 56–59. 10.1016/j.catcom.2016.11.016. [DOI] [Google Scholar]
  20. Shah D.; Kaur H. Resin-trapped gold nanoparticles: An efficient catalyst for reduction of nitro compounds and Suzuki-Miyaura coupling. J. Mol. Catal. A: Chem. 2014, 381, 70–76. 10.1016/j.molcata.2013.10.004. [DOI] [Google Scholar]
  21. Kumar S.; Sharma S.; Das P. Supported Gold Nanoparticles-Catalyzed Microwave-Assisted Hydration of Nitriles to Amides under Base-Free Conditions. Adv. Synth. Catal. 2016, 358, 2889–2894. 10.1002/adsc.201600199. [DOI] [Google Scholar]
  22. Karwa S. L.; Rajadhyaksha R. A. Selective catalytic hydrogenation of nitrobenzene to hydrazobenzene. Ind. Eng. Chem. Res. 1988, 27, 21–24. 10.1021/ie00073a005. [DOI] [Google Scholar]
  23. Shiraishi Y.; Katayama M.; Hashimoto M.; Hirai T. Photocatalytic hydrogenation of azobenzene to hydrazobenzene on cadmium sulfide under visible light irradiation. Chem. Commun. 2018, 54, 452–455. 10.1039/C7CC08428E. [DOI] [PubMed] [Google Scholar]
  24. Serna P.; Corma A. Transforming Nano Metal Nonselective Particulates into Chemoselective Catalysts for Hydrogenation of Substituted Nitrobenzenes. ACS Catal. 2015, 5, 7114–7121. 10.1021/acscatal.5b01846. [DOI] [Google Scholar]
  25. Gelder E. A.; Jackson S. D.; Lok C. M. The hydrogenation of nitrobenzene to aniline: a new mechanism. Chem. Commun. 2005, 4, 522–524. 10.1039/B411603H. [DOI] [PubMed] [Google Scholar]
  26. Corma A.; Concepción P.; Serna P. A different reaction pathway for the reduction of aromatic nitro compounds on gold catalysts. Angew. Chem., Int. Ed. 2007, 46, 7266–7269. 10.1002/anie.200700823. [DOI] [PubMed] [Google Scholar]
  27. Mahata A.; Rai R. K.; Choudhuri I.; Singh S. K.; Pathak B. Direct vs. indirect pathway for nitrobenzene reduction reaction on a Ni catalyst surface: a density functional study. Phys. Chem. Chem. Phys. 2014, 16, 26365–26374. 10.1039/C4CP04355C. [DOI] [PubMed] [Google Scholar]
  28. Liu Q.; Xu Y.; Qiu X.; Huang C.; Liu M. Chemoselective hydrogenation of nitrobenzenes activated with tuned Au/h-BN. J. Catal. 2019, 370, 55–60. 10.1016/j.jcat.2018.12.008. [DOI] [Google Scholar]
  29. Li H.-Q.; Liu X.; Zhang Q.; Li S.-S.; Liu Y.-M.; He H.-Y.; Cao Y. Deoxygenative coupling of nitroarenes for the synthesis of aromatic azo compounds with CO using supported gold catalysts. Chem. Commun. 2015, 51, 11217–11220. 10.1039/C5CC03134F. [DOI] [PubMed] [Google Scholar]
  30. Doherty S.; Knight J. G.; Backhouse T.; Summers R. J.; Abood E.; Simpson W.; Paget W.; Bourne R. A.; Chamberlain T. W.; Stones R.; Lovelock K. R. J.; Seymour J. K.; Isaacs M. A.; Hardacre C.; Daly H.; Rees N. H. Highly Selective and Solvent-Dependent Reduction of Nitrobenzene to N-Phenylhydroxylamine, Azoxybenzene and Aniline Catalyzed by Phosphino-Modified Polymer Immobilized Ionic Liquid-Stabilized AuNPs. ACS Catal. 2019, 9, 4777–4791. 10.1021/acscatal.9b00347. [DOI] [Google Scholar]
  31. Yu L.; Zhang Q.; Li S. S.; Huang J.; Liu Y. M.; He H. Y.; Cao Y. Gold-Catalyzed Reductive Transformation of Nitro Compounds Using Formic Acid: Mild, Efficient, and Versatile. ChemSusChem 2015, 8, 3029–3035. 10.1002/cssc.201500869. [DOI] [PubMed] [Google Scholar]
  32. Su Y.; Li X.; Wang Y.; Zhong H.; Wang R. Gold nanoparticles supported by imidazolium-based porous organic polymers for nitroarene reduction. Dalton Trans. 2016, 45, 16896–16903. 10.1039/C6DT03050E. [DOI] [PubMed] [Google Scholar]
  33. Shil A. K.; Das P. Solid supported platinum (0) nanoparticles catalyzed chemo-selective reduction of nitroarenes to N-arylhydroxylamines. Green Chem. 2013, 15, 3421–3428. 10.1039/c3gc41179f. [DOI] [Google Scholar]
  34. Bouillon M. E.; Meyer H. H. The 4.4′-benzidine rearrangement of 4-alkyl substituted hydrazobenzenes. Tetrahedron 2016, 72, 3151–3161. 10.1016/j.tet.2016.03.103. [DOI] [Google Scholar]
  35. Ling X.; Xiong Y.; Huang R.; Zhang X.; Zhang S.; Chen C. Synthesis of benzidine derivatives via FeCl3· 6H2O-promoted oxidative coupling of anilines. J. Org. Chem. 2013, 78, 5218–5226. 10.1021/jo4002504. [DOI] [PubMed] [Google Scholar]
  36. Kumar S.; Das P. Solid-supported ruthenium (0): an efficient heterogeneous catalyst for hydration of nitriles to amides under microwave irradiation. New J. Chem. 2013, 37, 2987–2990. 10.1039/c3nj00493g. [DOI] [Google Scholar]
  37. Shil A. K.; Kumar S.; Reddy C. B.; Dadhwal S.; Thakur V.; Das P. Supported palladium nanoparticle-catalyzed carboxylation of aryl halides, alkenylsilanes, and organoboronic acids employing oxalic acid as the C1 source. Org. Lett. 2015, 17, 5352–5355. 10.1021/acs.orglett.5b02701. [DOI] [PubMed] [Google Scholar]
  38. Guha N. R.; Sharma S.; Bhattacherjee D.; Thakur V.; Bharti R.; Reddy C. B.; Das P. Oxidative “reverse-esterification” of ethanol with benzyl/alkyl alcohols or aldehydes catalyzed by supported rhodium nanoparticles. Green Chem. 2016, 18, 1206–1211. 10.1039/C5GC02513C. [DOI] [Google Scholar]
  39. Bharti R.; Bal Reddy C.; Kumar S.; Das P. Supported palladium nanoparticle-catalysed Suzuki–Miyaura cross-coupling approach for synthesis of aminoarylbenzosuberene analogues from natural precursor. Appl Organometal Chem. 2017, 31, e3749 10.1002/aoc.3749. [DOI] [Google Scholar]
  40. Shibasaki-Kitakawa N.; Honda H.; Kuribayashi H.; Toda T.; Fukumura T.; Yonemoto T. Biodiesel production using anionic ion-exchange resin as heterogeneous catalyst. Bioresour. Technol. 2007, 98, 416–421. 10.1016/j.biortech.2005.12.010. [DOI] [PubMed] [Google Scholar]
  41. Bodamer G. W.; Kunin R. Behavior of Ion Exchange Resins in Solvents Other Than Water-Swelling and Exchange Characteristics. Ind. Eng. Chem. 1953, 45, 2577–2580. 10.1021/ie50527a057. [DOI] [Google Scholar]
  42. Bernar I.; Rutjes F. P.; Elemans J. A.; Nolte R. J. Aerobic Epoxidation of Low-Molecular-Weight and Polymeric Olefins by a Supramolecular Manganese Porphyrin Catalyst. Catalysts 2019, 9, 195. 10.3390/catal9020195. [DOI] [Google Scholar]
  43. Wunder S.; Polzer F.; Lu Y.; Mei Y.; Ballauff M. Kinetic analysis of catalytic reduction of 4-nitrophenol by metallic nanoparticles immobilized in spherical polyelectrolyte brushes. J. Phys. Chem. C 2010, 114, 8814–8820. 10.1021/jp101125j. [DOI] [Google Scholar]
  44. Aditya T.; Pal A.; Pal T. Nitroarene reduction: a trusted model reaction to test nanoparticle catalysts. Chem. Commun. 2015, 51, 9410–9431. 10.1039/C5CC01131K. [DOI] [PubMed] [Google Scholar]
  45. Layek K.; Kantam M. L.; Shirai M.; Nishio-Hamane D.; Sasaki T.; Maheswaran H. Gold nanoparticles stabilized on nanocrystalline magnesium oxide as an active catalyst for reduction of nitroarenes in aqueous medium at room temperature. Green Chem. 2012, 14, 3164–3174. 10.1039/c2gc35917k. [DOI] [Google Scholar]
  46. International Union of Pure and Applied Chemistry (IUPAC). Ionisation Constants of Organic Acids in Aqueous Solution, IUPAC Chemical Data Series No. 23; Serjeant E. P., Dempsey B., Eds.; Pergamon Press, Inc.: New York, 1979; p 989. [Google Scholar]
  47. Lange N. A.Lange’s Handbook of Chemistry, 13th ed.; Dean J. A., Ed.; McGraw-Hill Book Co.: New York, 1985; pp 5–51. [Google Scholar]
  48. For more information and data, see Supporting Information

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao0c00402_si_001.pdf (3.3MB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES