Next Article in Journal
MoO3-Based Photocatalysts for the Depolymerization of Lignin Under UV-Vis Light
Previous Article in Journal
High-Performance CoxNiy@NC/SiO2 Catalysts Derived from ZIF-67 for Enhanced Hydrogenation of 1-Nitronaphthalene
Previous Article in Special Issue
Effects of Zn Doping on the Morphology and H2 Production Activity of Truncated Octahedral Cu2O Photocatalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Innovative Copper-Based Heterogeneous Catalyst for Chan–Lam Cross-Coupling

1
Department of Physical Chemistry, Palacký University Olomouc, 77900 Olomouc, Czech Republic
2
Department of Organic Chemistry, Palacký University Olomouc, 77900 Olomouc, Czech Republic
*
Authors to whom correspondence should be addressed.
Catalysts 2026, 16(1), 94; https://doi.org/10.3390/catal16010094
Submission received: 28 November 2025 / Revised: 5 January 2026 / Accepted: 13 January 2026 / Published: 16 January 2026
(This article belongs to the Collection Nanotechnology in Catalysis)

Abstract

The synthesis, in particular the industrial production, of pharmaceuticals requires a broad arsenal of synthetic reactions capable of selectively forming specific structural motifs and assembling smaller building blocks into complex molecules. The Chan–Evans–Lam cross-coupling reaction, which forms a bond between a N-nucleophile and an aryl group from a boronic acid, catalysed by copper salts, is a typical example of this synthetic route. Considering the toxicity of copper and the stringent regulatory limits for its residues in final pharmaceutical products, a heterogeneous catalytic approach offers a viable alternative for this transformation. In this work, we present a simply and reproducibly synthesized catalyst based on copper nanoparticles supported on reduced graphene oxide (Cu-rGO), with high efficiency in a model Chan–Lam reaction involving benzimidazole and aniline derivatives with substituted boronic acids.

Graphical Abstract

1. Introduction

Despite the undeniable advantages of homogeneous catalysts used in organic synthesis, it is sometimes more appropriate to opt for heterogeneous catalysts, which are easily separable (via filtration, dialysis, or centrifugation). This feature becomes particularly important on a larger scale, both for obtaining a pure substance (especially in the pharmaceutical industry) and for the potential reuse of the catalyst. Since reactions with heterogeneous catalysts occur only on their surface, a crucial parameter is the surface area-to-mass (or volume) ratio. Nanomaterials exhibit significantly higher values of this parameter compared to standard heterogeneous catalysts, and therefore, they are frequently studied for catalytic applications. These nanomaterials can exist in a free dispersion or be anchored onto various types of supports, such as TiO2 [1,2,3], SiO2 [4,5,6,7], Fe3O4 [8,9], graphene, or carbon nitride. The support not only improves separability but may also enhance the stability of the nanoparticles themselves and can participate in the reaction or influence the optical, electrical, or magnetic properties of the resulting material.
Graphene, a two-dimensional carbon-based material discovered in 2004 [10], has found diverse applications in many fields from biomedicine [11,12,13,14] and electronics [15,16,17] to water purification technologies [18,19] and catalysis [20,21,22]. It also serves as a support for metal nanoparticles [23,24,25,26,27] or their compounds [28,29,30,31]. Graphene can be prepared via chemical exfoliation, such as the Hummers method [32,33,34], which starts from graphite and proceeds through graphene oxide, which is subsequently reduced to yield the final two-dimensional structure. The resulting material is more precisely referred to as reduced graphene oxide (rGO) [35,36,37,38].
The Chan–Evans–Lam cross-coupling, independently reported by its authors in 1998 [39,40,41], is a reaction that forms C–heteroatom bonds, typically involving amines [42,43,44,45] or nitrogen-containing heterocycles (especially imidazole and benzimidazole [46]), or hydroxyl groups [47,48,49] reacting with arylboronic acids or their esters [50,51,52]. Analogous reactions have also been described involving sulfur [53,54,55], selenium [56,57], or tellurium [58] atoms. The catalyst is typically copper-based. The general reaction scheme (Scheme 1) is as follows:
Due to its ability to form C–heteroatom bonds and thus connect two smaller structures, this reaction is a significant and potentially valuable tool, both for constructing specific structural motifs and for total synthesis of biologically active, particularly pharmaceutically relevant compounds. The Chan–Lam cross-coupling has been used, for example, in the synthesis of N-vinylcinnamaldehyde nitrones [59], N-arylphosphoramidates [60], N-arylcytisine derivatives [61], drugs based on N,N′-dimethylurea [62], aminobiphenylsulfonamides [63], and, in enantioconvergent form, chiral benzylamides [64]. Given their pharmaceutical applications, these syntheses are subject to strict limits on copper content in the final product, which often poses a problem for homogeneous catalysis.
In the early stages of applying heterogeneous catalysis to the Chan–Lam coupling, copper ions (CuII) bound in polymeric complexes [65,66] or metallic copper [67] were used. Over the following years, development progressed in the area of polymeric complexes [68,69] and zeolites [70], as well as nanoparticles, both as copper supports via organic linkers [71,72] and as catalysts themselves [73].
Today, a wide range of nanocomposites has been explored for heterogeneous catalysis of this reaction, including metallic copper nanoparticles anchored on g-C3N4 [74]. For unsupported nanoparticles, procedures using composite nanoparticles like Cu2O/Cu [75] or CuFe2O4 [76] (also usable in Sonogashira coupling) have been published. Some studies have also introduced photocatalysis for the Chan–Lam coupling [77], including the use of composite nanoparticles bound to graphene oxide: Cu2S:NiS2@C/rGO [78]. Beyond nanoparticles, various copper complexes [79,80,81] have also been immobilized on rGO for catalyzing the Chan–Lam cross-coupling.
The aim of the experimental work described in this article was to develop a simple and reproducible method for synthesizing a heterogeneous catalyst based on copper nanoparticles bound to reduced graphene oxide, and to study its applicability along with potential limitations for the Chan–Lam cross-coupling reaction, with a focus on aniline derivatives as substrates. The use of such a catalyst offers the advantage of minimizing copper contamination in the final product, which is a crucial factor, especially for pharmaceutical applications, due to copper’s toxicity [82,83,84,85].

2. Results and Discussion

All the chemicals used are commonly available for purchase.

2.1. Catalyst Synthesis

Graphene oxide (GO), used as an intermediate in the preparation of the catalyst, was synthesized using a modified Hummers method. This method is based on a multi-step, exothermic oxidation of graphite, followed by the separation of its layers [86] due to non-covalent repulsive interactions caused by the formation of electronegative functional groups.
In a round-bottom flask, graphite (1 g; Aldrich Chemistry; for synthesis; powder; <20 μm) was mixed with NaNO3 (0.5 g) and H2SO4 (23 mL). The mixture was stirred continuously while being cooled in an ice bath, and solid KMnO4 (3 g) was added very slowly. This was followed by heating and stirring the mixture (35 °C; 30 min). Then, demineralized water (46 mL) was added very slowly due to the accompanying exothermic reaction. The mixture was subsequently heated (98 °C; 30 min) and cooled to room temperature. After the addition of demineralized water (140 mL) and H2O2 (1 mL; 35%), the mixture was decanted, and the solid precipitate was washed using a centrifuge (3500 rpm), five times with water and once with ethanol. The product was dried in a vacuum oven (55 °C; 100 mbar) and ground into a fine powder, which was then used in subsequent steps.
The prepared GO (1 g) was then dispersed in 250 mL of H2O using an ultrasonic bath. A solution of Cu(NO3)2 (20 mL; 0,33 M) was then added, and the mixture was stirred for 1 h. Subsequently, the temperature was raised to 80 °C, and after the addition of N2H4 (2,7 mL; 50–60%), the mixture was stirred for an additional 3 h. After cooling to room temperature, the product was washed on a fritted filter with water and ethanol, dried in a vacuum oven (55 °C, 100 mbar), and used in the form of a fine powder.
The yield of the catalyst synthesized in this manner is quantitative (any losses are attributable to the decantation and washing procedures). The material appears as a fine well-filtrable black powder that is readily affected by static electricity. It is stable when stored at room temperature.

2.2. Catalyst Characterization

A sample of the prepared material (Cu-rGO) was dispersed in 2% nitric acid and placed in an ultrasonic bath for several minutes. The dispersion was then filtered through a microfilter, diluted with water, and analyzed using a ContrAA 300 atomic absorption spectrometer (Analytik Jena, Jena, Germany). The resulting copper mass fraction was 8%.
The material was further analyzed using electron microscopy, specifically transmission electron microscopes LVEM 5 (Delong Instruments, Brno, Czechia), JEOL 2010F (JEOL, Tokyo, Japan), JEOL 2010 H (JEOL, Tokyo, Japan), and high-resolution microscope HRTEM (Titan 2, Tokyo, Japan). TEM images clearly show that the final material (Figure 1C) consists of isolated spherical nanoparticles with diameters in the range of a few nanometers.
Additional measurements were carried out using a DXR Raman spectrometer (Thermo Scientific, Waltham, MA, USA) and an iS50 ATR infrared spectrometer (Nicolet, Waltham, WI, USA). Baseline correction was applied to all spectra.
The infrared (IR) spectrum reveals several prominent bands typical for C, H, and O bonds. This technique was primarily employed to characterize the composite and to verify the presence of functional groups before and after reduction. The most prominent band is a broad peak below 3500 cm−1, corresponding to O–H stretching vibrations. Another key region lies between 1800–1700 cm−1, indicating the presence of carbonyl groups (C=O), which gradually transitions into the ~1600 cm−1 region, where the C=C stretching vibrations dominate. Additional signals between approximately 1400–800 cm−1 can be attributed to σ-bond C–O vibrations [87]. A comparison of the spectra (Figure 2A) clearly shows that the transition from GO to rGO during synthesis is accompanied by a reduction of oxygen-containing functional groups, with the successful deposition of copper nanoparticles.
The Raman spectra of GO and rGO (Figure 2B) exhibit three characteristic bands, commonly referred to as the G, D, and 2D bands. The G band, observed around 1580–1600 cm−1, reflects the planar structure of the layer (i.e., the entire material in the case of graphene oxide) and arises from the sp2 hybridization of carbon atoms. It is also prominently present in graphite. The D band, around 1350 cm−1, indicates defects in the planar structure and is typically weaker in graphite. The 2D band is a higher-order overtone of the D band. Unlike the D band, it does not reflect defects, and its presence in the spectrum confirms the identity of graphene and provides information on its thickness, i.e., the number of layers [88].
To characterize the oxidation state of copper, X-ray diffraction (XRD) analysis was performed using a D8 ADVANCE diffractometer (Bruker, Billerica, MA, USA). The XRD pattern (Figure 3) supports all previous characterization results. The dominant peak at the low-angle region of the spectrum (2θ = 15°) in GO is a typical indicator of the material and can be used to calculate the interlayer spacing (~0.8 nm) between graphene sheets. A broad band around 25° corresponds to the functional reduction of GO to rGO and also confirms the retention of sp2 hybridization in the graphene lattice. Additional sharp peaks correspond to the crystal lattices of metallic copper (Cu0). All observed signals are consistent with the literature [89,90], though they appear slightly shifted by a few degrees due to the selected method and the reference used in the XRD measurements.
To characterize the material using X-ray photoelectron spectroscopy, Nexsa G2 XPS system (Thermo Fisher Scientific, Waltham, MA, USA) was used. In the spectrum (Figure 4), as expected, we observe peaks corresponding to carbon (280.3–298 eV), oxygen (252.02–545 eV), and residual nitrogen (394.29–410 eV). As for copper (926.6–960.55 eV), the profile of its bands indicates the presence of several different oxidation states—not only Cu(0), as seen in the XRD spectra, but also Cu(I) and Cu(II).
BET isotherm measurements for material characterization were performed using a 3Flex instrument, version 5.02 (Micromeritics, Norcross, GA, USA). The hysteresis between the adsorption and desorption isotherms (Figure 5) indicates the presence of pores, likely resulting from multilayered graphene structures. Such multilayering is, however, typical for chemically synthesized reduced graphene oxide (rGO). For single-point adsorption, the total pore volume of pores with a width less than 40.3122 nm at p/p° = 0.950 was 0.1419 cm3/g. The BET surface area was determined to be 77.55 ± 0.236 m2/g.
Summarizing the outcomes of the synthesis and subsequent characterization of composite catalyst Cu-rGO, it can be concluded that graphene oxide (GO) was successfully prepared using a modified Hummers’ method. Subsequent reduction by hydrazine together with Cu salt addition produces a final form of the catalyst with regularly dispersed Cu nanoparticles on the surface of rGO sheets.
The resulting material (Cu-rGO) contains 8% copper by mass, as determined by atomic absorption spectroscopy (AAS). Copper is present on the surface in the form of spherical nanoparticles with sizes in the low nanometer range. In addition to copper, the material is primarily composed of carbon, with minor amounts of nitrogen and oxygen. Oxygen is a typical component of rGO and contributes to the anchoring of metals or other linkers onto the carbon framework, making complete reduction of oxygen groups neither expected nor desirable.
The presence of nitrogen can be attributed to residuals from the synthesis process, where nitrogen appeared in several forms—as an oxidizing agent (NaNO3) during graphite oxidation, a copper precursor (Cu(NO3)2), and as a reducing agent (N2H4).
Copper is present predominantly in the Cu(0) oxidation state, though Cu(I) and Cu(II) species were also detected. The hysteresis loop in the adsorption isotherm indicates the presence of pores, with a pore volume close to 0.14 cm3/g, and a BET surface area of around 77.55 m2/g.

2.3. Chan–Lam Cross-Coupling

General Procedure and Monitoring

General procedure: A nitrogen nucleophile (0.526 mmol; 1 equiv.), arylboronic acid (1.16 mmol; 2.2 equiv.), and triethylamine (Et3N; 146 µL; 1.99 mmol; 2 equiv.) were dissolved in 1 mL of methanol (MeOH). After the addition of the Cu/rGO catalyst (20.2 mg; 0.05 equiv.), the resulting suspension was stirred for 24 h at 30 °C (30 °C was chosen as a temperature close to ambient, yet precisely standardized).
The reaction mixture was then filtered through diatomaceous earth and washed with MeOH. The filtrate was concentrated under reduced pressure and further purified by column chromatography on a silica gel column (40–63 μm particle size).
TLC and UPLC-MS were used to monitor the reaction progress and analyse the products.
Thin-layer chromatography (TLC) was performed using pre-coated silica gel plates. Detection was carried out under UV light (at 250 or 366 nm), followed by visualization using solutions of KMnO4, ninhydrin, or PMA (phosphomolybdic acid).
UPLC-MS analysis was performed using a system composed of an Acquity UPLC chromatograph equipped with a PDA and MS detector (Waters, Milford, MA, USA). A C18 column X-Select HSS T3 (2.5 μm, 3.0 mm × 50 mm, with a guard column, Waters, Milford, MA, USA) was used at 30 °C with a flow rate of 600 μL/min. The mobile phases (MP) were: (A) 0.01 M ammonium acetate in H2O, (B) acetonitrile (ACN). Elution was carried out using a gradient program, linearly programmed according to the specific method (normal, slow, etc.). The ESI source operated at a capillary voltage of 3 kV, with a desolvation temperature of 350 °C and a source temperature of 120 °C.

2.4. Results of Catalytic Experiments

To evaluate the applicability of the Cu-rGO catalyst, the Chan–Lam cross-coupling reaction was chosen, which involves coupling an aromatic boronic acid (tolylboronic acid) with a suitable N-nucleophile. For this study, benzimidazole [46,91,92,93,94] was chosen as the N-nucleophile.
A series of solvents was tested, among which alcohols, especially methanol and to a lesser extent butanol, proved to be suitable, as these are standard solvents for this reaction [95,96,97]. Among bases, the amines triethylamine (Et3N) and DIPEA (diisopropylethylamine [42,98,99]) performed best. Inorganic bases like K2CO3 or Cs2CO3 slowed the reaction compared to tertiary amines because the reaction mixture contained two chemically distinct heterogeneous phases.
The model reaction of benzimidazole with tolylboronic acid was carried out in the presence of triethylamine and 5 mol% of the Cu/rGO catalyst in methanol. After 24 h, UPLC-MS analysis showed complete conversion.
The influence of para-substituents on the reaction progress is summarized in Table 1. The data indicate that the introduction of electron-withdrawing groups (EWG) significantly decreases reactivity, while electron-donating groups (EDG) enhance it. Conversely, substitution at the 2-position of benzimidazole completely inhibits the reaction, likely due to a strong steric effect.
Regarding arylboronic acids, para-substitution generally did not pose a steric problem, so reactivity was primarily influenced by electronic effects. Electron-rich tolyl- (2b) and methoxyphenylboronic acids (2c) enhanced reactivity. We did not use 4-hydroxyphenylboronic acid due to chemoselectivity and protodeboronation issues [100]. The electron-withdrawing nitro group (2d) significantly reduced it and promoted the formation of side products. Phenylboronic acid (2a) gave a 51% yield.
As noted above, substitution at the 2-position of benzimidazole inhibits the reaction. Therefore, aniline derivatives were selected as suitable substrates for further investigation instead of benzimidazole. Each aniline derivative reacted similarly with four arylboronic acids. Again, two effects, electronic and steric, governed the reactivity of both reactants, were observed.
From the perspective of aniline and its derivatives (6), markedly different reactivities were observed depending on substitution:
Aniline (6a) reacted with all boronic acids (confirmed by UPLC-MS). Even with the challenging −NO2 substituted boronic acid (2d), a 25% isolated yield was obtained, the highest for this boronic acid. However, phenylboronic acid (2a) gave a mixture of side products.
With p-methylaniline (6b), which exerts a weak positive inductive effect, it was unclear if it reacted better or worse than unsubstituted aniline, as results varied with the boronic acid. The best results were with electron-donating p-methoxyphenylboronic acid (2c). The reaction with p-nitrophenylboronic acid produced only traces of product and many side products, so the product was not isolated.
p-Anisidine (6c) contains a methoxy group with a positive mesomeric effect and a sterically accessible amino group, supporting reactivity. Products were isolated with all boronic acids, with the highest yield from tolylboronic acid (2b; 53%).
3,4,5-Trimethoxyaniline (6d), an even more electron-rich substrate, gave a 70% yield with phenylboronic acid (2a). However, reactivity decreased with other boronic acids.
The methyl group in o-toluidine (6e) increases electron density but caused a dramatic drop in reactivity, likely due to steric hindrance at the ortho position limiting effective adsorption on the catalyst. The best result (12%) was with methoxyphenylboronic acid (2c). Two products from acids 2a and 2b with very low conversion were not isolated.
In contrast, meta-substitution with the same alkyl group (6f) did not cause steric hindrance but rather enhanced reactivity compared to ortho-substitution (6b), giving results comparable to para-methoxy substitution (6c). All products were isolated. A substrate with an electron-withdrawing and sterically demanding ortho-nitro group (6g) predictably gave minimal UPLC-MS response and no isolated products.
From these results (Table 2), it is clear that methyl substitution at para- or meta-positions does not slow the reaction. However, ortho-substitution causes reduced reactivity, likely due to steric hindrance limiting efficient adsorption of the amino group on the catalyst. The ortho-nitro group causes a pronounced synergistic negative effect: steric hindrance combined with electron withdrawal by the negative mesomeric effect severely suppresses the reaction.

3. Conclusions

This work described the preparation of a catalyst composed of copper nanoparticles anchored on reduced graphene oxide, containing 8% of copper (w/w). The prepared catalyst was successfully applied in cross-coupling synthesis, specifically in the Chan–Lam cross-coupling reaction. When applied to the reaction of substituted anilines with phenylboronic acids, it was demonstrated that, besides electronic effects, the steric accessibility of reactive sites on the substrate significantly influences reactivity. A clear example was toluidine, where the methyl group in the ortho-position slowed the reaction, while substitutions in the meta- and para-positions allowed the reaction to proceed similarly to unsubstituted aniline. This finding points to a limitation in the use of this catalyst material, but conversely, it provides a valuable predictive insight into how effective the reaction will be forgiven substrates. Another advantage is the very straightforward and reproducible synthesis of the presented catalyst material, as well as its easy separation from the reaction mixture. This enables the convenient isolation of pure products free from copper contamination—a crucial criterion, especially for industrial applications and opens doors for further research.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal16010094/s1. Figure S1, TEM images: (a) graphene oxide, (b) Cu-rGO composite, (c) same composite as (b), magnified view.

Author Contributions

Conceptualization, J.S., R.P., D.P., B.L., P.C. and L.K.; Methodology, D.P.; Validation, D.P. and P.C.; Investigation, J.S., D.P. and L.K.; Data curation, J.S., R.P., P.C. and Ľ.L.; Writing—original draft, J.S.; Writing—review & editing, R.P., D.P., B.L., P.C., L.K. and Ľ.L.; Visualization, J.S., R.P. and Ľ.L.; Supervision, R.P. and B.L.; Project administration, L.K. All authors have read and agreed to the published version of the manuscript.

Funding

The research presented in this article was supported by the following grants: IGA_PrF_2025_024 and IGA_PrF_2025_014.

Data Availability Statement

All relevant data discussed (e.g., in the form of graphs or tables) are presented directly in the research article/Supplementary Materials. Raw data are available upon request.

Acknowledgments

At this point, we would like to express our sincere gratitude to the following colleagues for their contributions to the article and for performing the measurements: Alexandra Rancová (BET), Barbora Štefková (Raman spectroscopy), Vít Procházka (XRD), Petr Martin (XPS), Ondřej Tomanec (HRTEM), Klára Čépe and Jana Stráská (TEM/SEM) and all from Palacký University Olomouc. We also wish to thank the entire institution for providing facilities and financial support, as well as the taxpayers of the Czech Republic. Dedicated to King Vratislav II.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of this study; in the collection, analyses, or interpretation of data; in the writing of this manuscript; or in the decision to publish the results.

References

  1. Chupas, P.J.; Chapman, K.W.; Jennings, G.; Lee, P.L.; Grey, C.P. Watching Nanoparticles Grow: The Mechanism and Kinetics for the Formation of TiO2-Supported Platinum Nanoparticles. J. Am. Chem. Soc. 2007, 129, 13822–13824. [Google Scholar] [CrossRef]
  2. Huang, Z.; Sun, S.; Ma, M.; Liu, Y.; Zhong, X.; Chen, J.; Gao, F.; Hai, G.; Huang, X. Facile Synthesis of TiO2 Supported Pd Nanoparticles for Efficient Photocatalytic CO2 Reduction to CH4 with H2O. Sustain. Mater. Technol. 2025, 43, e01247. [Google Scholar] [CrossRef]
  3. Serra-Pérez, E.; Dražić, G.; Takashima, M.; Ohtani, B.; Kovačič, S.; Žerjav, G.; Tušar, N.N. Influence of the Surface Structure of the TiO2 Support on the Properties of the Au/TiO2 Photocatalyst for Water Treatment under Visible Light. Catal. Today 2024, 437, 114764. [Google Scholar] [CrossRef]
  4. Ulusal, F. The Synthesis and Cytotoxicity Analysis of Gold Nanoparticles Coated with Mesoporous SiO2 from Gold Treatment Wastewater for Medical Applications. Inorg. Chem. Commun. 2024, 159, 111867. [Google Scholar] [CrossRef]
  5. Kim, T.S.; O’Connor, C.R.; Reece, C. Interrogating Site Dependent Kinetics over SiO2-Supported Pt Nanoparticles. Nat. Commun. 2024, 15, 2074. [Google Scholar] [CrossRef] [PubMed]
  6. Hoque, A.; Nawarathne, C.P.; Alvarez, N.T. Vertically Aligned Carbon Nanotubes from Premade Binary Metal Oxide Nanoparticles on Bare SiO2. Carbon 2025, 235, 120086. [Google Scholar] [CrossRef]
  7. Wang, F.; Tian, X.; Shi, Y.; Fan, W.; Liu, Q. Photocatalytic CO2 Methanation over the Ni/SiO2 Catalysts for Performance Enhancement. Int. J. Hydrogen Energy 2024, 68, 1382–1392. [Google Scholar] [CrossRef]
  8. Deng, J.; Li, F.; Qi, Z.; Huang, W.; Wan, Z.; Zhang, L.; Zheng, D.; Li, G.; Zhang, F. A Novel Redox Synergistic Mechanism of Peroxymonosulfate Activation Using Pd-Fe3O4 for Ultra-Fast Chlorinated Hydrocarbon Degradation. Appl. Catal. B Environ. Energy 2024, 359, 124499. [Google Scholar] [CrossRef]
  9. Payamifar, S.; Foroozandeh, A.; Abdouss, M.; Marjani, A.P. Magnetic Nickel Nanoparticle Catalyst on β-Cyclodextrin-Modified Fe3O4 for Nitroarene Hydrogenation. Sci. Rep. 2024, 14, 28493. [Google Scholar] [CrossRef]
  10. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric Field Effect in Atomically Thin Carbon Films Supplementary. Science 2004, 306, 666–669. [Google Scholar] [CrossRef]
  11. Derakhshi, M.; Daemi, S.; Shahini, P.; Habibzadeh, A.; Mostafavi, E.; Ashkarran, A.A. Two-Dimensional Nanomaterials beyond Graphene for Biomedical Applications. J. Funct. Biomater. 2022, 13, 27. [Google Scholar] [CrossRef]
  12. Kumar, R.; Singh, D.P.; Muñoz, R.; Amami, M.; Singh, R.K.; Singh, S.; Kumar, V. Graphene-Based Materials for Biotechnological and Biomedical Applications: Drug Delivery, Bioimaging and Biosensing. Mater. Today Chem. 2023, 33, 101750. [Google Scholar] [CrossRef]
  13. Kaur, H.; Garg, R.; Singh, S.; Jana, A.; Bathula, C.; Kim, H.S.; Kumbar, S.G.; Mittal, M. Progress and Challenges of Graphene and Its Congeners for Biomedical Applications. J. Mol. Liq. 2022, 368, 120703. [Google Scholar] [CrossRef] [PubMed]
  14. Zare, P.; Aleemardani, M.; Seifalian, A.; Bagher, Z.; Seifalian, A.M. Graphene Oxide: Opportunities and Challenges in Biomedicine. Nanomaterials 2021, 11, 1083. [Google Scholar] [CrossRef] [PubMed]
  15. He, W.; Zhai, L.; Yam, C.Y.; Xu, Z. Current Transients in Graphene Electronics under Single-Particle Irradiation. Small Sci. 2023, 3, 2300011. [Google Scholar] [CrossRef] [PubMed]
  16. Xiang, Z.; Wang, H.; Zhao, P.; Fa, X.; Wan, J.; Wang, Y.; Xu, C.; Yao, S.; Zhao, W.; Zhang, H.; et al. Hard Magnetic Graphene Nanocomposite for Multimodal, Reconfigurable Soft Electronics. Adv. Mater. 2024, 36, e2308575. [Google Scholar] [CrossRef]
  17. Banerjee, R.; Chowdhury, A.H.; Kumar, P.S.; Wang, C.; Goel, S.; Raj, P.M. Laser-Induced Graphene Supercapacitors on Flex Substrates for Package-Integrated Power Supply. In Proceedings of the 2023 4th International Symposium on 3D Power Electronics Integration and Manufacturing, 3D-PEIM, Miami, FL, USA, 1–3 February 2023. [Google Scholar]
  18. Pan, Y.; Shi, Z.; Li, J.; Zhang, Z.; Li, X.; Zhuang, Z.; Mo, Y.; Liang, J.; Wang, Z.; An, M.; et al. Graphene Oxide Laminates Intercalated with Prussian Blue Nanocube as a Photo-Fenton Self-Cleaning Membrane for Enhanced Water Purification. J. Memb. Sci. 2023, 672, 121465. [Google Scholar] [CrossRef]
  19. Sheikholeslami, M.; Khalili, Z.; Mousavi, S.J. Influence of Self-Cleaning Coating on Performance of Photovoltaic Solar System Utilizing Mixture of Phase Change Material and Graphene Nanoparticle. J. Build. Eng. 2023, 77, 107540. [Google Scholar] [CrossRef]
  20. Yan, Y.; Shin, W.I.; Chen, H.; Lee, S.M.; Manickam, S.; Hanson, S.; Zhao, H.; Lester, E.; Wu, T.; Pang, C.H. A Recent Trend: Application of Graphene in Catalysis. Carbon Lett. 2021, 31, 177–199. [Google Scholar] [CrossRef]
  21. Fan, F.R.; Wang, R.; Zhang, H.; Wu, W. Emerging Beyond-Graphene Elemental 2D Materials for Energy and Catalysis Applications. Chem. Soc. Rev. 2021, 50, 10983–11031. [Google Scholar] [CrossRef]
  22. Huang, L.; Cheng, L.; Ma, T.; Zhang, J.J.; Wu, H.; Su, J.; Song, Y.; Zhu, H.; Liu, Q.; Zhu, M.; et al. Direct Synthesis of Ammonia from Nitrate on Amorphous Graphene with Near 100% Efficiency. Adv. Mater. 2023, 35, e2211856. [Google Scholar] [CrossRef]
  23. Huang, L.W.; Jeng, H.T.; Su, W.B.; Chang, C.S. Indirect Interactions of Metal Nanoparticles through Graphene. Carbon 2021, 174, 132–137. [Google Scholar] [CrossRef]
  24. Choudhary, P.; Ramalingam, B.; Das, S.K. Rational Design of Antimicrobial Peptide Conjugated Graphene-Silver Nanoparticle Loaded Chitosan Wound Dressing. Int. J. Biol. Macromol. 2023, 246, 125347. [Google Scholar] [CrossRef]
  25. Zhou, S.; Zeng, X.; Yan, X.; Xie, F.; Fahlman, B.D.; Wang, C.; Li, W. High Aspect Ratio Copper Nanowires and Copper Nanoparticles Decorated by Reduced Graphene Oxide for Flexible Transparent Conductive Electrodes. Appl. Surf. Sci. 2022, 604, 154597. [Google Scholar] [CrossRef]
  26. Mehmandoust, M.; Erk, N.; Karaman, O.; Karimi, F.; Bijad, M.; Karaman, C. Three-Dimensional Porous Reduced Graphene Oxide Decorated with Carbon Quantum Dots and Platinum Nanoparticles for Highly Selective Determination of Azo Dye Compound Tartrazine. Food Chem. Toxicol. 2021, 158, 112698. [Google Scholar] [CrossRef]
  27. Kostoglou, N.; Liao, C.W.; Wang, C.Y.; Kondo, J.N.; Tampaxis, C.; Steriotis, T.; Giannakopoulos, K.; Kontos, A.G.; Hinder, S.; Baker, M.; et al. Effect of Pt Nanoparticle Decoration on the H2 Storage Performance of Plasma-Derived Nanoporous Graphene. Carbon 2021, 171, 294–305. [Google Scholar] [CrossRef]
  28. Purabgola, A.; Mayilswamy, N.; Kandasubramanian, B. Graphene-Based TiO2 Composites for Photocatalysis & Environmental Remediation: Synthesis and Progress. Environ. Sci. Pollut. Res. 2022, 29, 32305–32325. [Google Scholar]
  29. Park, S.; Choi, S.H.; Kim, J.M.; Ji, S.; Kang, S.; Yim, S.; Myung, S.; Kim, S.K.; Lee, S.S.; An, K.S. Nanoarchitectonics of MXene Derived TiO2/Graphene with Vertical Alignment for Achieving the Enhanced Supercapacitor Performance. Small 2024, 20, e2305311. [Google Scholar] [CrossRef]
  30. Gonçalves, D.A.; Alencar, L.M.; Anjos, J.P.B.; Orth, E.S.; Souza, V.H.R. Ultrasensitive Electrochemical Detection of Glyphosate Using Crumpled Graphene/Copper Oxide Nanocomposite. Mater. Today Commun. 2023, 36, 106746. [Google Scholar] [CrossRef]
  31. Tu, X.; Xie, Y.; Gao, F.; Ma, X.; Lin, X.; Huang, X.; Qu, F.; Ping, L.; Yu, Y.; Lu, L. Self-Template Synthesis of Flower-like Hierarchical Graphene/Copper Oxide@copper(II) Metal-Organic Framework Composite for the Voltammetric Determination of Caffeic Acid. Microchim. Acta 2020, 187, 258. [Google Scholar] [CrossRef]
  32. Marcano, D.C.; Kosynkin, D.V.; Berlin, J.M.; Sinitskii, A.; Sun, Z.; Slesarev, A.; Alemany, L.B.; Lu, W.; Tour, J.M. Improved Synthesis of Graphene Oxide. ACS Nano 2010, 4, 4806–4814. [Google Scholar] [CrossRef]
  33. Mondal, P.; Sinha, A.; Salam, N.; Roy, A.S.; Jana, N.R.; Islam, S.M. Enhanced Catalytic Performance by Copper Nanoparticle-Graphene Based Composite. RSC Adv. 2013, 3, 5615–5623. [Google Scholar] [CrossRef]
  34. Hummers, W.S.; Offeman, R.E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  35. Smith, A.T.; LaChance, A.M.; Zeng, S.; Liu, B.; Sun, L. Synthesis, Properties, and Applications of Graphene Oxide/Reduced Graphene Oxide and Their Nanocomposites. Nano Mater. Sci. 2019, 1, 31–47. [Google Scholar] [CrossRef]
  36. Tarcan, R.; Todor-Boer, O.; Petrovai, I.; Leordean, C.; Astilean, S.; Botiz, I. Reduced Graphene Oxide Today. J. Mater. Chem. C 2020, 8, 1198–1224. [Google Scholar] [CrossRef]
  37. Kiranakumar, H.V.; Thejas, R.; Naveen, C.S.; Khan, M.I.; Prasanna, G.D.; Reddy, S.; Oreijah, M.; Guedri, K.; Bafakeeh, O.T.; Jameel, M. A Review on Electrical and Gas-Sensing Properties of Reduced Graphene Oxide-Metal Oxide Nanocomposites. Biomass Convers. Biorefin. 2024, 14, 12625–12635. [Google Scholar] [CrossRef]
  38. Shabil Sha, M.; Anwar, H.; Musthafa, F.N.; Al-Lohedan, H.; Alfarwati, S.; Rajabathar, J.R.; Khalid Alahmad, J.; Cabibihan, J.J.; Karnan, M.; Kumar Sadasivuni, K. Photocatalytic Degradation of Organic Dyes Using Reduced Graphene Oxide (RGO). Sci. Rep. 2024, 14, 3608. [Google Scholar] [CrossRef] [PubMed]
  39. Evans, D.A.; Katz, J.L.; West, T.R. ChemInform Abstract: Synthesis of Diaryl Ethers Through the Copper-Promoted Arylation of Phenols with Arylboronic Acids. An Expedient Synthesis of Thyroxine. ChemInform 1998, 29. [Google Scholar] [CrossRef]
  40. Chan, D.M.T.; Monaco, K.L.; Wang, R.P.; Winters, M.P. New N- and O-Arylations with Phenylboronic Acids and Cupric Acetate. Tetrahedron Lett. 1998, 39, 2933–2936. [Google Scholar] [CrossRef]
  41. Lam, P.Y.S.; Clark, C.G.; Saubern, S.; Adams, J.; Winters, M.P.; Chan, D.M.T.; Combs, A. New Aryl/Heteroaryl C-N Bond Cross-Coupling Reactions via Arylboronic Acid/Cupric Acetate Arylation. Tetrahedron Lett. 1998, 39, 2941–2944. [Google Scholar] [CrossRef]
  42. West, M.J.; Fyfe, J.W.B.; Vantourout, J.C.; Watson, A.J.B. Mechanistic Development and Recent Applications of the Chan-Lam Amination. Chem. Rev. 2019, 119, 12491–12523. [Google Scholar] [CrossRef]
  43. Robinson, H.; Oatley, S.A.; Rowedder, J.E.; Slade, P.; Macdonald, S.J.F.; Argent, S.P.; Hirst, J.D.; McInally, T.; Moody, C.J. Late-Stage Functionalization by Chan–Lam Amination: Rapid Access to Potent and Selective Integrin Inhibitors. Chem.—A Eur. J. 2020, 26, 7678–7684. [Google Scholar] [CrossRef]
  44. Kuliukhina, D.S.; Malysheva, A.S.; Averin, A.D.; Savelyev, E.N.; Orlinson, B.S.; Novakov, I.A.; Beletskaya, I.P. Chan–Lam N-Arylation of Adamantane-Containing Amines. Russ. J. Org. Chem. 2023, 59, 2107–2116. [Google Scholar] [CrossRef]
  45. Han, Y.; Zhang, M.; Zhang, Y.Q.; Zhang, Z.H. Copper Immobilized at a Covalent Organic Framework: An Efficient and Recyclable Heterogeneous Catalyst for the Chan-Lam Coupling Reaction of Aryl Boronic Acids and Amines. Green Chem. 2018, 20, 4891–4900. [Google Scholar] [CrossRef]
  46. Janíková, K.; Jedinák, L.; Volná, T.; Cankař, P. Chan-Lam Cross-Coupling Reaction Based on the Cu2S/TMEDA System. Tetrahedron 2018, 74, 606–617. [Google Scholar] [CrossRef]
  47. Halford-McGuff, J.M.; Israel, E.M.; West, M.J.; Vantourout, J.C.; Watson, A.J.B. Direct Chan–Lam Amination and Etherification of Aryl BMIDA Reagents. Eur. J. Org. Chem. 2022, 2022, e202200993. [Google Scholar] [CrossRef]
  48. Wu, M. Chan-Lam Oxygen-Cyclopropylation of Aliphatic Alcohols; Smith College: Northampton, MA, USA, 2022. [Google Scholar]
  49. Derosa, J.; O’Duill, M.L.; Holcomb, M.; Boulous, M.N.; Patman, R.L.; Wang, F.; Tran-Dubé, M.; McAlpine, I.; Engle, K.M. Copper-Catalyzed Chan-Lam Cyclopropylation of Phenols and Azaheterocycles. J. Org. Chem. 2018, 83, 3417–3425. [Google Scholar] [CrossRef] [PubMed]
  50. Naya, L.; Larrosa, M.; Rodríguez, R.; Cruces, J. Selective Copper-Promoted Cross-Coupling Reaction of Anilines and Alkylboranes. Tetrahedron Lett. 2012, 53, 769–772. [Google Scholar] [CrossRef]
  51. Vantourout, J.C.; Miras, H.N.; Isidro-Llobet, A.; Sproules, S.; Watson, A.J.B. Spectroscopic Studies of the Chan-Lam Amination: A Mechanism-Inspired Solution to Boronic Ester Reactivity. J. Am. Chem. Soc. 2017, 139, 4769–4779. [Google Scholar] [CrossRef]
  52. Grayson, J.D.; Dennis, F.M.; Robertson, C.C.; Partridge, B.M. Chan-Lam Amination of Secondary and Tertiary Benzylic Boronic Esters. J. Org. Chem. 2021, 86, 9883–9897. [Google Scholar] [CrossRef] [PubMed]
  53. Bhowmik, A.; Yadav, M.; Fernandes, R.A. Room Temperature Nickel-Catalyzed Cross-Coupling of Aryl-Boronic Acids with Thiophenols: Synthesis of Diarylsulfides. Org. Biomol. Chem. 2020, 18, 2447–2458. [Google Scholar] [CrossRef]
  54. Lam, L.Y.; Ma, C. Chan-Lam-Type C-S Coupling Reaction by Sodium Aryl Sulfinates and Organoboron Compounds. Org. Lett. 2021, 23, 6164–6168. [Google Scholar] [CrossRef]
  55. Greenwood, N.S.; Ellman, J.A. Sulfur-Arylation of Sulfenamides via Chan-Lam Coupling with Boronic Acids: Access to High Oxidation State Sulfur Pharmacophores. Org. Lett. 2023, 25, 2830–2834. [Google Scholar] [CrossRef]
  56. Yu, J.T.; Guo, H.; Yi, Y.; Fei, H.; Jiang, Y. The Chan-Lam Reaction of Chalcogen Elements Leading to Aryl Chalcogenides. Adv. Synth. Catal. 2014, 356, 749–752. [Google Scholar] [CrossRef]
  57. Sun, N.; Zheng, K.; Zhang, M.; Zheng, G.; Jin, L.; Hu, B.; Shen, Z.; Hu, X. Cu-Catalysed Chan-Lam Synthesis of Unsymmetrical Aryl Chalcogenides under Aqueous Micellar Conditions. Green Chem. 2023, 25, 2782–2789. [Google Scholar] [CrossRef]
  58. Fernandes, R.A.; Bhowmik, A.; Yadav, S.S. Advances in Cu and Ni-Catalyzed Chan–Lam-Type Coupling: Synthesis of Diarylchalcogenides, Ar2–X (X = S, Se, Te). Org. Biomol. Chem. 2020, 18, 9583–9600. [Google Scholar] [CrossRef]
  59. Chen, H.; Wei, C.; Pang, G.L.; Liang, C.; Mo, D.L.; Ma, X.P. Synthesis of N-Vinyl Cinnamaldehyde Nitrones through Atropisomeric Quinoxaline-Derived N, N, O-Ligand-Promoted Chan-Lam Reaction. Org. Lett. 2022, 24, 6013–6017. [Google Scholar] [CrossRef]
  60. Dangroo, N.A.; Ara, T.; Dar, B.A.; Khuroo, M.A. Copper Catalyzed Tandem Chan–Lam Type C—N and Staudinger-Phosphite N—P Coupling for the Synthesis of N-Arylphosphoramidates. Catal. Commun. 2019, 118, 76–80. [Google Scholar] [CrossRef]
  61. Sánchez-Velasco, O.A.; Saavedra-Olavarría, J.; Araya-Santelices, D.A.A.; Hermosilla-Ibáñez, P.; Cassels, B.K.; Pérez, E.G. Synthesis OfN-Arylcytisine Derivatives Using the Copper-Catalyzed Chan-Lam Coupling. J. Nat. Prod. 2021, 84, 1985–1992. [Google Scholar] [CrossRef]
  62. Saikia, R.; Das, S.; Almin, A.; Mahanta, A.; Sarma, B.; Thakur, A.J.; Bora, U. N,N′-Dimethylurea as an Efficient Ligand for the Synthesis of Pharma-Relevant Motifs through Chan-Lam Cross-Coupling Strategy. Org. Biomol. Chem. 2023, 21, 3143–3155. [Google Scholar] [CrossRef]
  63. Sharif, M.; Mahmood, A.; Kanwal, A.; Ahmad, M.; Rasool, N.; Usman, M.Q.; Al Mughram, M.H.; Al-Harerhi, A.A.M.; Shah, T.A.; Skaikh, I.I. Facile Synthesis of Aminobiphenyl Sulfonamides via Chan–Lam Coupling and Their Biological Evaluation as Potent Carbonic Anhydrase Inhibitors. Sci. Rep. 2025, 15, 25661. [Google Scholar] [CrossRef]
  64. Vu, J.; Haug, G.C.; Schubert, T.J.; Head, J.F.; Patton, R.S.; Dong, Y. Enantioconvergent Chan–Lam Coupling: Synthesis of Chiral Benzylic Amides via Cu-Catalyzed Deborylative Amidation. J. Am. Chem. Soc. 2025, 147, 25527–25535. [Google Scholar] [CrossRef] [PubMed]
  65. Anuradha; Kumari, S.; Pathak, D.D. Synthesis and Development of Chitosan Anchored Copper(II) Schiff Base Complexes as Heterogeneous Catalysts for N-Arylation of Amines. Tetrahedron Lett. 2015, 56, 4135–4142. [Google Scholar] [CrossRef]
  66. Wang, B.; Yang, P.; Ge, Z.W.; Li, C.P. A Porous Metal–Organic Framework as Active Catalyst for Multiple C–N/C–C Bond Formation Reactions. Inorg. Chem. Commun. 2015, 61, 13–15. [Google Scholar] [CrossRef]
  67. Bao, J.; Tranmer, G.K. The Solid Copper-Mediated C–N Cross-Coupling of Phenylboronic Acids under Continuous Flow Conditions. Tetrahedron Lett. 2016, 57, 654–657. [Google Scholar] [CrossRef]
  68. Khatioda, R.; Pathak, D.; Sarma, B. Cu(II) Complex onto a Pyridine-Based Porous Organic Polymer as a Heterogeneous Catalyst for Nitroarene Reduction. ChemistrySelect 2018, 3, 6309–6320. [Google Scholar] [CrossRef]
  69. Khosravi, A.; Mokhtari, J.; Naimi-Jamal, M.R.; Tahmasebi, S.; Panahi, L. Cu2(BDC)2(BPY)-MOF: An Efficient and Reusable Heterogeneous Catalyst for the Aerobic Chan-Lam Coupling Prepared via Ball-Milling Strategy. RSC Adv. 2017, 7, 46022–46027. [Google Scholar] [CrossRef]
  70. Garnier, T.; Sakly, R.; Danel, M.; Chassaing, S.; Pale, P. Chan-Lam-Type C-N Cross-Coupling Reactions under Base- and Ligand-Free CuI-Zeolite Catalysis. Synthesis 2017, 49, 1223–1230. [Google Scholar] [CrossRef]
  71. Sharma, H.; Mahajan, H.; Jamwal, B.; Paul, S. Cu@Fe3O4-TiO2-L-Dopa: A Novel and Magnetic Catalyst for the Chan-Lam Cross-Coupling Reaction in Ligand Free Conditions. Catal. Commun. 2018, 107, 68–73. [Google Scholar] [CrossRef]
  72. Dutta, M.M.; Phukan, P. Cu-Doped CoFe2O4 Nanoparticles as Magnetically Recoverable Catalyst for CN Cross-Coupling Reaction. Catal. Commun. 2018, 109, 38–42. [Google Scholar] [CrossRef]
  73. Sharma, N.; Choudhary, A.; Kaur, M.; Sharma, C.; Paul, S.; Gupta, M. Modified Graphene Supported Ag-Cu NPs with Enhanced Bimetallic Synergistic Effect in Oxidation and Chan-Lam Coupling Reactions. RSC Adv. 2020, 10, 30048–30061. [Google Scholar] [CrossRef] [PubMed]
  74. Seyedi, N.; Shahabi Nejad, M.; Saidi, K.; Sheibani, H. Fabrication of Nitrogen-Enriched Graphene Oxide/Cu NPs as a Highly Efficient and Recyclable Heterogeneous Nanocatalyst for the Chan–Lam Cross-Coupling Reaction. Appl. Organomet. Chem. 2020, 34, e5307. [Google Scholar] [CrossRef]
  75. Sarmah, M.; Sarmah, D.; Dewan, A.; Bora, P.; Boruah, P.K.; Das, M.R.; Bharali, P.; Bora, U. Dual Responsive Sustainable Cu2O/Cu Nanocatalyst for Sonogashira and Chan-Lam Cross-Coupling Reactions. Catal. Lett. 2023, 153, 1423–1437. [Google Scholar] [CrossRef]
  76. Banda, P.G.; Kurup, G.B.; Mucherla, R. Magnetically Recoverable CuFe2O4 Nanocatalyst: Dual Catalytic Action in Sonogashira and Chan-Lam Coupling Reactions. J. Organomet. Chem. 2024, 1016, 123259. [Google Scholar] [CrossRef]
  77. Di, J.Q.; Zhang, M.; Chen, Y.X.; Wang, J.X.; Geng, S.S.; Tang, J.Q.; Zhang, Z.H. Copper Anchored on Phosphorus G-C3N4as a Highly Efficient Photocatalyst for the Synthesis OfN-Arylpyridin-2-Amines. Green Chem. 2021, 23, 1041–1049. [Google Scholar] [CrossRef]
  78. Yusuf, M.; Hira, S.A.; Lim, H.; Song, S.; Park, S.; Park, K.H. Core-Shell Cu2S:NiS2@C Hybrid Nanostructure Derived from a Metal-Organic Framework with Graphene Oxide for Photocatalytic Synthesis of N-Substituted Derivatives. J. Mater. Chem. A 2021, 9, 9018–9027. [Google Scholar] [CrossRef]
  79. Mirza-Aghayan, M.; Saeedi, M.; Boukherroub, R. Carbon–Nitrogen Bond Formation Using Modified Graphene Oxide Derivatives Decorated with Copper Complexes and Nanoparticles. Appl. Organomet. Chem. 2021, 35, e6327. [Google Scholar] [CrossRef]
  80. Mittal, A.; Kumari, S.; Parmanand; Yadav, D.; Sharma, S.K. A New Copper Complex on Graphene Oxide: A Heterogeneous Catalyst for N-Arylation and C-H Activation. Appl. Organomet. Chem. 2020, 34, e5362. [Google Scholar] [CrossRef]
  81. Kumar, A.; Layek, S.; Agrahari, B.; Kujur, S.; Pathak, D.D. Graphene Oxide Immobilized Copper(II) Schiff Base Complex [GO@AF-SB-Cu]: A Versatile Catalyst for Chan-Lam Coupling Reaction. ChemistrySelect 2019, 4, 1337–1345. [Google Scholar] [CrossRef]
  82. Wang, Y.; Yan, Q.; Shi, Y.; Long, M. Copper Toxicity in Animals: A Review. Biol. Trace Elem. Res. 2024, 203, 2675–2686. [Google Scholar] [CrossRef] [PubMed]
  83. Gaetke, L.M.; Chow, C.K. Copper Toxicity, Oxidative Stress, and Antioxidant Nutrients. Toxicology 2003, 189, 147–163. [Google Scholar] [CrossRef]
  84. Charkiewicz, A.E. Is Copper Still Safe for Us? What Do We Know and What Arethe Latest Literature Statements? Curr. Issues Mol. Biol. 2024, 46, 8441–8463. [Google Scholar] [CrossRef]
  85. Royer, A.; Sharman, T. Copper Toxicity; StatPearls Publishing: Treasure Island, FL, USA, 2025. [Google Scholar]
  86. Chen, X.; Qu, Z.; Liu, Z.; Ren, G. Mechanism of Oxidization of Graphite to Graphene Oxide by the Hummers Method. ACS Omega 2022, 7, 23503–23510. [Google Scholar] [CrossRef] [PubMed]
  87. Brusko, V.; Khannanov, A.; Rakhmatullin, A.; Dimiev, A.M. Unraveling the Infrared Spectrum of Graphene Oxide. Carbon 2024, 229, 119507. [Google Scholar] [CrossRef]
  88. Ossonon, B.D.; Bélanger, D. Synthesis and Characterization of Sulfophenyl-Functionalized Reduced Graphene Oxide Sheets. RSC Adv. 2017, 7, 27224–27234. [Google Scholar] [CrossRef]
  89. Nasrollahzadeh, M.; Babaei, F.; Fakhri, P.; Jaleh, B. Synthesis, Characterization, Structural, Optical Properties and Catalytic Activity of Reduced Graphene Oxide/Copper Nanocomposites. RSC Adv. 2015, 5, 10782–10789. [Google Scholar] [CrossRef]
  90. Elzey, S.; Baltrusaitis, J.; Bian, S.; Grassian, V.H. Formation of Paratacamite Nanomaterials via the Conversion of Aged and Oxidized Copper Nanoparticles in Hydrochloric Acidic Media. J. Mater. Chem. 2011, 21, 3162–3169. [Google Scholar] [CrossRef]
  91. Györke, G.; Dancsó, A.; Volk, B.; Milen, M. Chan−Lam Arylation of Benzimidazole and Its Derivatives in the Presence of Copper-Containing Minerals. ChemistrySelect 2021, 6, 802–807. [Google Scholar] [CrossRef]
  92. Anupama, T.S.A.; Monica, V.; Małecki, J.G.; Keri, R.S.; Azam, M.; Al-Resayes, S.I.; Budagumpi, S. Chan−Evans−Lam Coupling for the Synthesis of N–Aryl Derivatives Catalyzed by Copper(I) Chloride and Sterically Varied Imidazolium Salts at Mild Reaction Conditions. J. Mol. Struct. 2024, 1301, 137362. [Google Scholar] [CrossRef]
  93. Zhou, X.; Yang, J.; Hao, Z.; Han, Z.; Lin, J.; Lu, G.L. Copper Complexes with N,N,N-Tridentate Quinolinyl Anilido-Imine Ligands: Synthesis and Their Catalytic Application in Chan−Lam Reactions. Molecules 2023, 28, 7406. [Google Scholar] [CrossRef]
  94. Ishida, M.; Adachi, R.; Kobayashi, K.; Yamamoto, Y.; Kawahara, C.; Yamada, T.; Aoyama, H.; Kanomata, K.; Akai, S.; Lam, P.Y.S.; et al. First Atroposelective Chan-Lam Coupling for the Synthesis of C-N Linked Biaryls. Chem. Commun. 2023, 60, 678–681. [Google Scholar] [CrossRef] [PubMed]
  95. Lan, J.B.; Chen, L.; Yu, X.Q.; You, J.S.; Xie, R.G. A Simple Copper Salt Catalysed the Coupling of Imidazole with Arylboronic Acids in Protic Solvent. Chem. Commun. 2004, 4, 188–189. [Google Scholar] [CrossRef] [PubMed]
  96. Clerc, A.; Bénéteau, V.; Pale, P.; Chassaing, S. Chan-Lam-Type Azidation and One-Pot CuAAC under CuI-Zeolite Catalysis. ChemCatChem 2020, 12, 2060–2065. [Google Scholar] [CrossRef]
  97. Pharande, P.S.; Mhaldar, P.M.; Lohar, T.R.; Ghotekar, S.K.; Chhowala, T.N.; Rashinkar, G.S.; Pore, D.M. A Selective Heterogeneous Cellulose Supported Schiff Base Cu(II) Catalyst for Chan–Evans–Lam Coupling. Res. Chem. Intermed. 2023, 49, 4541–4560. [Google Scholar] [CrossRef]
  98. Roy, T.; Mondal, K.; Halder, P.; Sengupta, A.; Das, P. CuF2/DTBP-Catalyzed Chan-Lam Coupling of Oxazolidinones with Arylboronic Acid Pinacol Ester: Scope and Application. J. Org. Chem. 2025, 90, 6219–6232. [Google Scholar] [CrossRef]
  99. Valencia, J.; Sánchez-Velasco, O.A.; Saavedra-Olavarría, J.; Hermosilla-Ibáñez, P.; Pérez, E.G.; Insuasty, D. N-Arylation of 3-Formylquinolin-2(1H)-Ones Using Copper(II)-Catalyzed Chan–Lam Coupling. Molecules 2022, 27, 8345. [Google Scholar] [CrossRef] [PubMed]
  100. Tomanová, M.; Jedinák, L.; Košař, J.; Kvapil, L.; Hradil, P.; Cankař, P. Synthesis of 4-Substituted Pyrazole-3,5-Diamines: Via Suzuki-Miyaura Coupling and Iron-Catalyzed Reduction. Org. Biomol. Chem. 2017, 15, 10200–10211. [Google Scholar] [CrossRef]
Scheme 1. General scheme of the Chan–Lam cross-coupling reaction.
Scheme 1. General scheme of the Chan–Lam cross-coupling reaction.
Catalysts 16 00094 sch001
Figure 1. HRTEM photos of Cu-rGO: (A) HRTEM, 1 μm, (B) HRTEM, 1 μm, element map, (C) HRTEM, 100 nm, (D) samples by EDS.
Figure 1. HRTEM photos of Cu-rGO: (A) HRTEM, 1 μm, (B) HRTEM, 1 μm, element map, (C) HRTEM, 100 nm, (D) samples by EDS.
Catalysts 16 00094 g001
Figure 2. (A) ATR-FTIR spectra of GO, rGO, and Cu-rGO, (B) Raman spectra of GO, rGO, and Cu-rGO (The patterns were shifted along the Y-axis to improve their visibility).
Figure 2. (A) ATR-FTIR spectra of GO, rGO, and Cu-rGO, (B) Raman spectra of GO, rGO, and Cu-rGO (The patterns were shifted along the Y-axis to improve their visibility).
Catalysts 16 00094 g002
Figure 3. XRD patterns of GO, rGO a Cu-rGO (The patterns were shifted along the Y-axis to improve their visibility).
Figure 3. XRD patterns of GO, rGO a Cu-rGO (The patterns were shifted along the Y-axis to improve their visibility).
Catalysts 16 00094 g003
Figure 4. XPS pattern of Cu-rGO.
Figure 4. XPS pattern of Cu-rGO.
Catalysts 16 00094 g004
Figure 5. Isotherm Plot of Cu-rGO.
Figure 5. Isotherm Plot of Cu-rGO.
Catalysts 16 00094 g005
Table 1. Isolated yields of Chan–Lam cross-coupling of benzimidazole with para-substituted phenylboronic acids products.
Table 1. Isolated yields of Chan–Lam cross-coupling of benzimidazole with para-substituted phenylboronic acids products.
Catalysts 16 00094 i001
Catalysts 16 00094 i002Isolated Yield (%)
Catalysts 16 00094 i0032a51
Catalysts 16 00094 i0042b24
Catalysts 16 00094 i0052c83
Catalysts 16 00094 i0062d8
Table 2. Isolated yields of Chan–Lam cross-coupling of aniline derivatives with para-substituted phenylboronic acids.
Table 2. Isolated yields of Chan–Lam cross-coupling of aniline derivatives with para-substituted phenylboronic acids.
Catalysts 16 00094 i007
Aniline DerivativeCatalysts 16 00094 i008Catalysts 16 00094 i009Catalysts 16 00094 i010Catalysts 16 00094 i011
Catalysts 16 00094 i0127461925
Catalysts 16 00094 i01317526
Catalysts 16 00094 i01412526
Catalysts 16 00094 i01570157
Catalysts 16 00094 i016126
Catalysts 16 00094 i01713325321
Catalysts 16 00094 i018
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Stehlík, J.; Pocklanová, R.; Profous, D.; Lapčíková, B.; Cankař, P.; Kvítek, L.; Lapčík, Ľ. Innovative Copper-Based Heterogeneous Catalyst for Chan–Lam Cross-Coupling. Catalysts 2026, 16, 94. https://doi.org/10.3390/catal16010094

AMA Style

Stehlík J, Pocklanová R, Profous D, Lapčíková B, Cankař P, Kvítek L, Lapčík Ľ. Innovative Copper-Based Heterogeneous Catalyst for Chan–Lam Cross-Coupling. Catalysts. 2026; 16(1):94. https://doi.org/10.3390/catal16010094

Chicago/Turabian Style

Stehlík, Jan, Radka Pocklanová, David Profous, Barbora Lapčíková, Petr Cankař, Libor Kvítek, and Ľubomír Lapčík. 2026. "Innovative Copper-Based Heterogeneous Catalyst for Chan–Lam Cross-Coupling" Catalysts 16, no. 1: 94. https://doi.org/10.3390/catal16010094

APA Style

Stehlík, J., Pocklanová, R., Profous, D., Lapčíková, B., Cankař, P., Kvítek, L., & Lapčík, Ľ. (2026). Innovative Copper-Based Heterogeneous Catalyst for Chan–Lam Cross-Coupling. Catalysts, 16(1), 94. https://doi.org/10.3390/catal16010094

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop