Apoorva
Shetty
ab,
Braja Gopal
Bag
c,
Uraiwan
Sirimahachai
d and
Gurumurthy
Hegde
*ab
aDepartment of Chemistry, Christ University, Hosur Road, Bangalore 560029, India
bCentre for Advanced Research and Development (CARD), Christ University, Hosur Road, Bangalore 560029, India. E-mail: murthyhegde@gmail.com
cDepartment of Chemistry & Chemical Technology, Vidyasagar University, Midnapore 721102, West Bengal, India
dDepartment of Chemistry and Centre of Excellence for Innovation in Chemistry, Faculty of Science, Prince of Songkla University, Hat Yai, Songkla 90112, Thailand
First published on 23rd January 2025
In this work, the synthesis and catalytic activity of thiol-functionalized copper-deposited porous carbon derived from dry oil palm leaves (Cu/TF-CNS) was investigated for the reduction of aromatic nitro compounds. The procedure to synthesize porous carbon nanospheres involves the pyrolysis of oil palm leaves in a nitrogen atmosphere at 1000 °C. The resulting porous carbon material was further functionalized with thiol groups to facilitate the uniform deposition of copper nanoparticles and serve as an efficient support. Excellent catalytic performance was shown by the Cu/TF-CNS catalyst in reducing aromatic nitro compounds to their corresponding aromatic amines with a low copper loading of only 4 mol% which is an inexpensive non-noble metal in the presence of NaBH4 as a reducing agent and EtOH/H2O as green solvents. The products were identified using 1H NMR spectroscopy. The catalyst was isolated from the reaction mixture and reused upto 10 cycles without any significant loss in the activity. The ICPAES analysis confirmed the successful incorporation of approximately 8.9% Cu during the deposition process and the reusability of the catalyst underscores its efficacy as a sustainable and effective heterogeneous catalyst for nitroarene reduction.
The catalytic reduction of aromatic nitro compounds remains attractive due to its widespread use in synthesizing intermediates and final products across various sectors such as dyes, agricultural chemicals, pharmaceuticals, and materials, both in laboratory and industrial settings. Despite numerous reported synthetic routes for aniline preparation from corresponding aromatic nitro compounds, there persists a need for the development of cost-effective catalysts with high activity.
The conventional method of producing amines involves reacting free hydrogen gas with the nitro groups while using either heterogeneous or homogeneous catalysts. However, because of the severe reaction conditions, including high temperatures and pressures this conventional procedure poses a higher risk.3–5 To address safety concerns, alternative hydrogen sources such as hydrazines, formic acid, alcohol derivatives, including glycerol, isopropanol, and ethanol have garnered interest. Although these alternatives are safer, environmentally friendly, and require less stoichiometric amounts, their hydrogenation process tends to be slow and occasionally demands elevated temperatures.6 Thus, sodium borohydride (NaBH4) is regarded as the best option for reducing R–NO2 to R–NH2, acting as a possible source of hydrogen.7 Sodium borohydride hydrolysis is a useful source of hydrogen due to its high volumetric density (10.8 wt% hydrogen content), excellent stability, and safety.3 The conventional use of sodium borohydride typically involves noble metal activators such as silver, platinum, or palladium.8,9
The strategy of employing metals as catalysts in organic reactions has a century-old tradition. Metallic nanoparticles, distinguished by unique physico-chemical properties attributed to a higher surface atom count, have garnered significant attention.10,11 Recently, noble metal-free alternatives have emerged, aiming to advance more sustainable chemical processes.8,12
Copper nanoparticles (CuNPs) are particularly appealing due to their abundant natural availability and cost-effectiveness. The simple and productive methods of preparing these Cu-based nanomaterials add to their attractiveness.13 Copper, as a versatile transition metal, can exist in various oxidation states (Cu0, CuI, CuII, and CuIII), enabling diverse reactions through both one- and two-electron pathways. The unique characteristics of Cu-based nanocatalysts find applications in nanotechnology, including catalytic organic transformations, electrocatalysis, and photocatalysis.13,14
In reduction reactions specifically, the active site of Cu exhibits strong affinity for hydrogen atoms and other molecules, which prolongs the lifespan of reactive species and increases the potential for reactions.15 However, to address potential issues such as aggregation in aqueous media, a promising approach involves the development of solid supports for the immobilization and stabilization of nanoparticles.10 Over time, the research emphasis has also been towards the development of supports, which is often more cost-effective and environmentally sustainable compared to focusing on active species. This aspect renders the field highly appealing with substantial potential. A fundamental approach to altering support properties involves manipulating their size and shape, achieved through downsizing into nanoparticles. This modification significantly impacts the resulting catalytic properties, enabling precise adjustments for the preferential exposure of specific active sites. Such modifications aim to maximize the availability of active sites for reagents, enhance overall activity, and guide the reaction towards the formation of a desired product, thereby achieving greater selectivity.16,17
In heterogeneous catalysis, porous solids serve as supports that facilitate the even dispersion of active metal or metal oxide catalysts while also helping to lower the overall cost of the catalyst.18 A wide range of support materials are available, including amorphous oxides, nanoporous crystalline oxides (zeolites), mesoporous oxides, and others.19 However, carbon based materials have found numerous applications in heterogeneous catalysis because of their abundance and low cost.20 Among these, nanocarbons stand out due to their numerous advantages, such as large surface area, high adsorption capacity, remarkable stability in aggressive media (acidic or basic), excellent thermal and mechanical stability, outstanding electronic properties, tunable porosity, and the ability to effectively anchor and disperse active metal nanoparticles.19,21 These properties enable nanocarbons to perform versatile functions. Hence significant research efforts continue to explore and unravel the fascinating chemistry of nanocarbon materials.22
The utilization of low-cost waste materials, such as lignocellulose,23 bagasse,24,25 and chitosan,26,27 for the synthesis of such carbonaceous supports/materials is imperative. These materials, characterized by both economic viability and environmental benignity, play a crucial role as raw materials in facilitating advanced catalytic transformations.
Using carbon as a support for heterogeneous catalysts presents significant challenges due to its complex surface characteristics and the potential for surface modification through various treatments.28 Introducing heteroatoms can create defects that modify the surface structure, providing anchoring points for metal nanoparticles or active sites.29,30 Chemical functionalization of nanocarbons is thus a valuable approach for introducing uniformly distributed anchoring points.31 Additionally, modified carbons often exhibit reduced aggregation compared to their pristine forms, resulting in improved dispersibility and enhanced catalytic performance.32 Sulfur-containing functional groups, particularly thiol (–SH), have attracted special attention due to their strong affinity for transition metals, making them effective for anchoring metal nanoparticles.33,34
In this study, we investigated the potential of using a low loading (4 mol%) of affordable Cu to develop a biomass derived thiol functionalized catalyst Cu/TF-CNS with good efficiency, yield, stability and recyclability for nitro amine reduction reactions. The porous carbons were synthesized from dry oil palm leaves at 1000 °C through a pyrolysis technique providing an efficient carbon support. Thiol functional groups were introduced onto the carbon via a chemical functionalization technique, and copper nanoparticles were then immobilized onto the porous nanocarbons, resulting in the formation of the Cu/TF-CNS catalyst. The effectiveness of Cu/TF-CNS was tested under different reaction conditions using the model reduction of 4-nitroaniline. Additionally, its versatility and recyclability were examined to determine its practical potential as an efficient heterogeneous catalyst. This study aims to deepen our understanding of the potential of using low-cost metals like copper effectively, as well as the impact of functionalization on the stability, recyclability and catalytic performance, thereby providing promising implications for developing cost-effective and sustainable materials.
![]() | ||
Scheme 1 Schematic representation for the synthesis of OPL-CNS followed by the synthesis of Cu/TF-CNS. |
About 100 mL of ACN was added to the activated acid functionalized carbon nanospheres (AF-CNS) taken in a 500 mL round bottom flask. To this, about 3 mL of DIPEA and 3 mL of 2-mercaptoethanol were added and stirred at RT for 3 h. The reaction mixture was then heated at 65 °C for 12 h. Upon the completion of the reaction, it was cooled to RT, and the solid was allowed to settle and washed with about 50 mL of water followed by 50 mL of ethanol, dried in a hot air oven at 50 °C for about 8 h to obtain thiol functionalized carbon nanospheres (TF-CNS). (Scheme 1, step 3).
The Fig. 1a–c illustrate the X-ray diffraction (XRD) patterns, FESEM image with the EDS plot and TEM image of the synthesized OPL-CNS. The XRD pattern of OPL-CNS exhibited two broad peaks around 24° and 44°, corresponding to the (002) and (100) lattice planes of graphitic carbon, (JCDPS card no. 75–1621) respectively.37 The broad nature of these peaks suggested the formation of an amorphous type of carbon material. The structural characteristics of OPL-CNS were initially analyzed using FESEM, as shown in Fig. 1b, where the CNS displayed a spherical morphology with an average diameter of 62 nm and the EDS analysis of OPL-CNS indicated that carbon accounted for approximately 79% of the composition. The TEM image of OPL-CNS Fig. 1c further confirmed the spherical morphology and revealed an uneven, flake-like structure, potentially contributing to the tendency of the material to form spherical aggregates.38
![]() | ||
Fig. 1 (a) X-ray diffraction pattern of OPL-CNS, (b) FESEM image along with the EDS plot of OPL-CNS and (c) TEM image of OPL-CNS. |
The crystallinity and morphological analysis was performed on the synthesized Cu/TF-CNS catalyst. The XRD pattern of Cu/TF-CNS Fig. 2a confirmed the presence of Cu particles on the carbon substrate. Along the broad peak at approximately 24°, indicative of the (002) facet of amorphous carbon, two sharp peaks were observed at 2θ = 42.4 and 73.7 corresponding to the (111) and (220) planes of fcc metallic Cu (JCPDS no. 04-0836).39 Additionally, two distinct sharp peaks at 2θ = 36.5 and 61.6 were attributed to the (111) and (220) planes of cubic phases of Cu2O, (JCPDS no. 78-2076) respectively.40 The XRD pattern showed that both Cu and Cu2O coexist in the material and the formation of Cu2O may be due to the nature of copper readily oxidizing in the presence of air. Fig. 2b presents the FESEM image and EDS profile of Cu/TF-CNS, illustrating the presence of copper particles deposited on the mesoporous carbon, while the carbon has retained its spherical morphology even after the deposition of copper nanoparticles. The EDS profile of the catalytic material Cu/TF-CNS gave the elemental percentages of carbon, oxygen, sulphur and copper to be 63%, 23.8%, 4.5%, and 8.6% respectively. The elemental percentage of copper on TF-CNS, as indicated by the EDS profile, aligns with the amount of copper deposited during the synthesis procedure and is consistent with the ICP-AES analysis data (Table S3, ESI†). The elemental mapping of the Cu/TF-CNS further demonstrated a uniform deposition of copper as well as the uniform functionalization of the carbon surface (Fig. 2d). The TEM image (Fig. 2c) of Cu/TF-CNS further confirms the deposition of copper on TF-CNS; the copper nanoparticles appeared as dark spots on the surface of the carbon.
XPS measurements were conducted to analyze the elemental composition and chemical bonding states of OPL-CNS, and the catalyst Cu/TF-CNS. These measurements provided insights into the oxidation state of the copper species associated with the catalyst under investigation and helped assess the role of copper species that contribute towards catalytic activity. The representative XPS survey spectra for OPL-CNS revealed the presence of C 1s and O 1s peaks, with mass concentrations of approximately 75% for C and 15% for O. Trace amounts of Ca (∼3%) and Si (∼6%) were also detected in the sample. The well deconvoluted C 1s spectrum (Fig. 3a) consisted of four distinct component peaks corresponding to sp3 C–C (285 eV), C–O (286 eV), C–O–C (286.8 eV), and CO (287.9 eV). Similarly, the O 1s spectrum (Fig. 3b) was deconvoluted into four individual component peaks corresponding to ether C–O–C (531.1 eV), carbonyl oxygen O–C
O (532.19 eV), non-carbonyl oxygen in O
C–O (533.1 eV), and hydroxy O–H (534.4 eV).38,41 The representative XPS survey spectra for Cu/TF-CNS revealed the presence of C 1s, O 1s, S 2p and Cu 2p peaks. The mass concentrations were found to be 61.8%, 20.3%, 1.8% and 9% for C, O, S, and Cu respectively. The binding energy (BE) of C 1s electrons in the Cu/TF-CNS catalyst was similar to that of OPL-CNS. The well deconvoluted C 1s spectrum (Fig. 3c) contained five distinct component peaks corresponding to sp3 C–C (285 eV), C–O (285.8 eV), C–O–C (286.8 eV), C
O (287.7 eV) and O–C
O (289.3 eV).42 However, due to the close binding energy values, the C 1s core spectrum cannot be precisely used to differentiate between the C–O, C–S and the desired C–SH bond. The O 1s spectrum (Fig. 3d) was deconvoluted into four individual component peaks corresponding to ether C–O–C (531.3 eV), carbonyl oxygen O–C
O (532.2 eV), non-carbonyl oxygen in O
C–O (533.3 eV), and hydroxy O–H (534.4 eV). Notably, there was an increase in percentage area for carbonyl oxygen O–C
O (532.2 eV) as compared to O 1s spectra of OPL-CNS, which may be attributed to the functionalization of the material resulting in an increase in carbonyl oxygen content. The S 2p spectrum (Fig. 3e) is also deconvoluted to reveal two peaks of C–S–H (S 2p) appearing at binding energies of 163.9 eV and 165.1 eV for Cu/TF-CNS,43 confirming the successful grafting of S–H onto the carbon surface. Additional peaks at 168.7 eV and169.5 eV can be attributed to S–O/O–S–O, likely resulting from the formal oxidation of sulfur during analysis.44 In the XPS spectra (Fig. 3f), two Cu peaks were observed at binding energies of approximately 932.9 eV and 952.8 eV, corresponding to Cu 2p3/2 and Cu 2p1/2, respectively.45 The presence of CuO (934.5 and 955.3 eV) indicates that copper on the surface can be easily oxidized at room temperature.
![]() | ||
Fig. 3 XPS spectrum: OPL-CNS (a and b) for C 1s and O 1s, respectively, and Cu/TF-CNS (c–f) for C 1s, O 1s, S 2p and Cu 2p, respectively. |
Additionally, peaks for Cu(OH)2 (935.5 and 954.4 eV) were also detected.46,47 The binding energies (BE) of Cu 2p3/2 and Cu 2p1/2 electrons in Cu/TF-CNS suggest the presence of both metallic Cu0 and Cu2+ species. The limited energy resolution of X-ray photoelectron spectroscopy (XPS) makes it challenging to differentiate between Cu and Cu2O since the difference in their 2p3/2 energies is less than the resolution of the XPS tool.48 These findings support the efficiency and recyclability of the catalyst in reduction reactions, as the successful functionalization of the support has led to better interaction between the support and the metallic particles, contributing to their effective recyclability and catalytic properties.
SL. no | Solvent | Catalyst loading (wt%) | Hydrogen source | Temperature | Time | Product conversion |
---|---|---|---|---|---|---|
Reaction conditions: 4-nitroaniline (1.0 mmol), Cu/TF-CNS, hydrogen source (5 mmol), solvent 3 mL.a Product conversion based on GCMS analysis.b Isolated yield. | ||||||
1 | EtOH/H2O | 20% | NaBH4 | 80 °C | 50 min | 58%a |
2 | EtOH | 20% | NaBH4 | 80 °C | 50 min | 15%a |
3 | H 2 O | 20% | NaBH 4 | 80 °C | 20 min | >99% |
4 | H2O | 5% | NaBH4 | 80 °C | 90 min | 12%a |
5 | H2O | 10% | NaBH4 | 80 °C | 50 min | >99%b |
6 | H2O | 15% | NaBH4 | 80 °C | 25 min | >99%b |
7 | H2O | 25% | NaBH4 | 80 °C | 20 min | >99%b |
8 | H2O | 20% | HCOOH | 80 °C | 120 min | 2%a |
9 | H2O | 20% | N2H4·2H2O | 80 °C | 120 min | No reaction |
10 | H2O | 20% | NaBH4 | RT | 150 min | 9%b |
11 | H2O | 20% | NaBH4 | 50 °C | 150 min | 90%b |
12 | H2O | 20% (Cu/OPL-CNS) | NaBH4 | 80 °C | 60 min | 96%b |
13 | H2O | 20% (OPL-CNS) | NaBH4 | 80 °C | 60 min | Trace |
14 | H2O | No catalyst | NaBH4 | 80 °C | 120 min | <5% |
15 | H2O | 20% | No H2 source | 80 °C | 120 min | No reaction |
The optimization studies were performed for the model reaction of the reduction of 4-nitroaniline. The selection of solvents played a crucial role. The aim was to use water as a green solvent since organic solvents tend to cause carcinogenic and genotoxic diseases; however there were some water insoluble starting materials, therefore the EtOH/H2O system was chosen to facilitate the solubility of such materials. Hence, the choice of solvent was evaluated keeping the hydrogen source (NaBH4), the catalyst loading (20 wt%) and the temperature (80 °C) constant. The experimental results for the reduction of 4-nitroaniline are summarized in Table 1 (entries 1–3), demonstrating promising conversion rates (>99%) for the H2O system in 20 min, whereas the EtOH system exhibited inadequate conversion, and the EtOH/H2O solvent system did not yield favorable outcome as compared to the H2O system in facilitating the reduction reaction; this may be due to the poor solubility of the starting material. Hence due to better conversion rate of the H2O system, this solvent system was optimized for further investigations.
Further investigations were performed to optimize the catalyst loading. Cu loadings 5 wt%, 10 wt%, 15 wt%, 20 wt% and 25 wt% were considered (Table 1, entries 3–7), keeping the solvent system (H2O), the hydrogen source (NaBH4) and the temperature (80 °C) constant. After examining the reactions, a 20 wt% loading (4 mol%) of Cu was selected for further studies due to its better product conversion within a shorter reaction time. Additionally, increasing the catalyst loading did not significantly affect the catalytic performance in this study.
In the next phase, the catalytic activity of Cu/TF-CNS was evaluated using NaBH4, formic acid and hydrazine hydrate as different hydrogen sources. Notably, the reaction using NaBH4 as a reducing agent exhibited significantly better conversion rates and yields compared to those using HCOOH and N2H4·2H2O which did not yield the desired results (Table 1, entries 3, 8 & 9). This enhanced performance may be attributed to the greater amount of hydride formed and transferred when using NaBH4 in conjunction with the Cu/TF-CNS catalyst under the optimized reaction conditions. Further optimization of the reaction temperature at RT, 50 °C and 80 °C showed that performing the reaction at 80 °C resulted in high conversion rates (Table 1, entries 3, 10 &11). Studies performed to examine the effect of functionalization of the catalyst showed that OPL-CNS (Table 1, entry 12) did not exhibit significant catalytic activity, while Cu/OPL-CNS (Table 1, entry 13), the non-thiol-functionalized catalyst, demonstrated reduced catalytic activity and required longer reaction times as compared to Cu/TF-CNS, the thiol-functionalized catalyst. The results confirmed that chemical functionalization, which introduces heteroatoms, creates uniformly distributed anchoring points for metal nanoparticles due to their stronger affinity for transition metals. This leads to improved dispersibility and enhanced catalytic performance. Subsequently, the role of the Cu/TF-CNS catalyst and the significance of the hydrogen source (NaBH4) were examined. In the absence of a catalyst, the reaction progressed slowly, with a conversion rate of less than 5% even after 120 min (Table 1, entry 14), while in the absence of a hydrogen source (NaBH4) the reaction did not occur at all (Table 1, entry 15). These results clearly demonstrate the crucial roles of both the Cu/TF-CNS catalyst and the hydrogen source in facilitating the efficient progress of the reduction reaction.
Entry | Substrate | Product (1P to 13P) | Reaction time | Isolated yield |
---|---|---|---|---|
Reaction conditions: aromatic nitro compound (1.0 mmol), NaBH4 (5 mmol), Cu/TF-CNS (20 wt%) temperature 80 °C, H2O (3 mL).a Solvent: 3 mL, EtOH/H2O (2![]() ![]() |
||||
1. |
![]() |
![]() |
180 min | 96% |
2. |
![]() |
![]() |
20 min | 98% |
3. |
![]() |
![]() |
90 mina | 91% |
4. |
![]() |
![]() |
45 min | 85% |
5. |
![]() |
![]() |
30 min | 62% |
6. |
![]() |
![]() |
35 mina | 73% |
7. |
![]() |
![]() |
30 mina | 79% |
8. |
![]() |
![]() |
35 mina | 85% |
9. |
![]() |
![]() |
80 min | 78% |
10. |
![]() |
![]() |
45 mina | 93% |
11. |
![]() |
![]() |
45 min | 87% |
12. |
![]() |
![]() |
25 min | 95% |
13. |
![]() |
![]() |
60 mina | 93% |
An examination of the morphology of the recycled catalyst using FESEM (Fig. 5b) revealed no significant changes. A hot filtration test was performed to further validate the heterogeneous nature of Cu/TF-CNS. The test demonstrated an approximately 50% conversion rate after 10 minutes, at which point the catalyst was removed from the reaction mixture. The reaction was then allowed to proceed for an additional 150 minutes without the catalyst, and TLC analysis after 2.5 h showed no significant increase in the conversion rate of the desired product. These results affirm the structural stability and heterogeneity of the catalyst.
Table 3 presents a comparative analysis of the catalytic activity of the Cu/TF-CNS catalyst with other previously reported noble and non-noble metal based heterogeneous catalysts for the reduction of 4-nitroaniline to benzene-1,4-diamine. From Table 3, it is evident that Cu/TF-CNS demonstrated favorable results regarding yield and reaction time compared to other catalytic systems. Although some systems achieved high yields, they often required larger quantities of more expensive metals, complex synthesis procedures or relatively longer reaction times. Additionally, the reusable Cu/TF-CNS catalyst offers several advantages including (i) the use of a cost effective, non-noble metal, (ii) high efficiency, excellent recyclability and environmental friendliness due to its operation in green solvents, (iii) enhanced safety, since it does not necessitate any specialized hydrogenation system, and (iv) ease of operation under ambient conditions.
Entry | Catalyst | Metal loading | Solvent | Hydrogen source | Temp | Time | Yield (%) | Ref. |
---|---|---|---|---|---|---|---|---|
1. | Pd–gCN | 5 mol% | EtOH | N2H4·H2O | 70 °C | 4 h | 97% | 49 |
2. | Pd@CTF | 1 mol% | EtOH/H2O | NH4COOH | 25 °C | 20 min | >99% | 50 |
3. | NiNPs/DNA | 2 mol% | H2O | NaBH4 | RT | 3 h | >99% | 51 |
4. | Ni@N-CNT | 2.4 wt% | MeOH | H2 (5bar) | 80 °C | 1.5 h | >99% | 52 |
5. | CuFe2O4@SiO2@g-C3N4/Cu | 35 wt% | H2O | NaBH4 | 55 °C | 20 min | 98% | 53 |
6. | CuO@C | 5 mol% | EtOH | NaBH4 | RT | 2 h | 86%% | 54 |
7. | Cu/TF-CNS | 4 mol% (20 wt%) | H2O | NaBH4 | 80 °C | 20 min | 98% | Present work |
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4cy01368a |
‡ 1H NMR and 13C NMR data of compounds. ICPAES data for the catalyst (ESI†). |
This journal is © The Royal Society of Chemistry 2025 |