Next Article in Journal
Opoka—Sediment Rock as New Type of Hybrid Mineral Filler for Polymer Composites
Previous Article in Journal
Chemometric Differentiation of Pistachios (Pistacia vera, Greek ‘Aegina’ Variety) from Two Different Harvest Years Using FTIR Spectroscopy and DRIFTS and Disk Techniques
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Carbon–Heteroatom Bond Formation via Coupling Reactions Performed on a Magnetic Nanoparticle Bed

by
Mahmood Tajbakhsh
1,*,
Ali Ramezani
1,
Mohammad Qandalee
2,
Mobina Falahati
1,
Carlos J. Durán-Valle
3,
Silvia Izquierdo
4 and
Ignacio M. López-Coca
4,*
1
Department of Organic Chemistry, Faculty of Chemistry, University of Mazandaran, Babolsar 47416-95447, Iran
2
Department of Basic Sciences, Garmsar Branch, Islamic Azad University, Garmsar 35816-31167, Iran
3
IACYS, Faculty of Sciences, University of Extremadura, 06006 Badajoz, Spain
4
INTERRA, School of Technology, University of Extremadura, 10003 Cáceres, Spain
*
Authors to whom correspondence should be addressed.
AppliedChem 2021, 1(2), 75-89; https://doi.org/10.3390/appliedchem1020007
Submission received: 23 July 2021 / Revised: 14 September 2021 / Accepted: 16 September 2021 / Published: 4 October 2021

Abstract

:
Cross-coupling reactions leading to carbon–heteroatom bonds yield compounds that attract substantial interest due to their role as structural units in many synthetic protocols for bioactive and natural products. Therefore, many research works aim at the improvement of heterogeneous catalytic protocols. We have studied the use of magnetite nanoparticles and solid base compounds in organic synthetic reactions in carbon–heteroatom bond formation because they can be flocculated and dispersed, and reversibly controlled by applying a magnetic field. In this work, we have developed an efficient and simple synthetic approach for the C–O/C–N cross-coupling reaction under ligand-free conditions by using CuI as a catalyst and KF/Fe3O4 as a base. We performed the nucleophilic aromatic substitution of electron-deficient aryl halides and phenols. It was found that both the solvent nature and the base have a profound influence on the reaction process. This approach affords good to excellent yields of arylated products. KF/Fe3O4 displayed convenient magnetic properties and could be easily separated from the reaction using a magnet and recycled several times without significant loss of catalytic activity. This method has been successfully investigated for the Ullmann coupling reaction.

Graphical Abstract

1. Introduction

Copper chemistry is extremely important because it can form Cu0, CuI, CuII, and CuIII oxidation states allowing one-electron or two-electron exchange. The different oxidation states can make useful interactions with different functional groups as Lewis acid or π-coordination. These features show remarkable activities allowing copper to catalyze different reactions [1].
Copper-complex catalytic materials have been successfully applied to cross-coupling reactions leading to carbon–heteroatom bonds. The chemicals thereby obtained attract substantial interest due to their important role as structural units in many synthetic protocols en route to bioactive and natural products [2,3,4,5].
Nucleophilic aromatic substitution (SNAr) reaction is an important methodology for the preparation of diaryl ethers from activated aryl halides or phenols [6,7]; for example, the coupling of activated aryl halides with different phenolic substrates in the presence of potassium fluoride/Clinoptilolite (KF/CP) was studied by Hosseini et al. [8]. Compared to typical transition-metal-mediated processes for the formation of a carbon–heteroatom bond, the SNAr reaction was found to be simple, showing mild reaction conditions, and was environmentally more convenient. An interesting alternative to this methodology is the coupling reaction of O, N, and S-arylation with 4-nitrochlorobenzene [9]. Recently, many researchers have aimed at the improvement of heterogeneous catalysts such as alumina-supported KOH, KF, and NaOH, transition-metal oxides [10,11,12], clay minerals or related materials [13,14], and zeolite-based catalysts [15,16]. Our interest in organic synthetic reactions developed by magnetite nanoparticle conditions prompted us to study the application of solid base compounds in carbon–heteroatom bond formation because they can be flocculated and dispersed and reversibly controlled by applying a magnetic field. By removing the external magnetic field, superparamagnetic nanoparticles can be dispersed completely in the reaction media, creating a vast surface that can be easily accessed by the substrate. Because of the high surface area and high catalytic activity, the practical utility of magnetic nanocomposites as catalysts has increased [17,18,19,20,21,22,23,24]. Nanomagnetic catalysts have been applied and studied widely in photocatalysts [25], biocatalysts [26], and phase-transfer catalysts [27]. Zhang and co-workers used nanoparticle Fe3O4-encapsulated CuO as a heterogeneous catalyst in the synthesis of diaryl ethers by the cross-coupling reaction; they also found that Fe3O4 alone did not show any catalytic effect [28]. Hu and co-workers reported a synthesis of the nanomagnetic catalyst KF/CaO–Fe3O4 by a facile impregnation method. The base catalyst KF/CaO–Fe3O4 was used to catalyze the transesterification of Stillingia oil and methanol for biodiesel production [29]. Using this method, we tried to prepare a nanomagnetic solid base catalyst KF/Fe3O4 that has an average particle diameter of ca. 20 nm. Other examples of the use of nanocatalysts in carbon-oxygen cross-coupling reactions for the synthesis of diaryl ethers can be found elsewhere [30]. Magnetic nanoparticles can selectively adsorb K ions and other alkali metal cations [29,31,32]. In the course of our previously published works for constructing of C–O/C–N bonds, we observed that KF impregnated on Fe3O4 (KF/Fe3O4) is a particularly useful and effective base for coupling reactions [8,33,34,35,36,37]. In a previous work, we found that the KF/Al2O3 combination provides a viable alternative to bases such as Cs2CO3, in the N-arylation of diazoles and the copper-catalyzed N-arylation of arylsulfonamides [33,38]. Here, the KF/Fe3O4 blend, which has the advantage of being separable with a magnet, along with an inexpensive Cu salt, can act satisfactorily in the C–O/C–N bond-formation reactions. The use of KF/Fe3O4 and CuX provides a synthesis with high selectivity in coupling reactions. We believe that KF supported on a magnetic Fe3O4 bed constitutes a good complement to other bases in copper-catalyzed coupling reactions. To the best of our knowledge, this is the first report on the use of KF/Fe3O4 in the C–N bond formation in the arylation of amides, carbazoles, and indoles and the synthesis of diaryl ethers.

2. Materials and Methods

2.1. Characterization

1H and 13C NMR spectra were recorded in Bruker Avance III spectrometer operating at 400.13 and 75.47 MHz, respectively; relevant spectra are available in the Supplementary Materials. Powder X-ray diffraction (XRD) was carried out on a Bruker ADVANCE-BRUKER D8 Discover (Cu K_ = 0/15,406 nm). The scanning rate was 1 min−1 in the 2Θ? range from 20 to 80°. Scanning electron microscopy (SEM) was recorded on a Hitachi S-1460 electron microscope, using an accelerating voltage of 200 kV. Melting points were determined using an Electrothermal IA 9100 Digital Melting Point apparatus. Thin-layer chromatography (TLC) was performed using 60 mesh silica gel Merck TLC plates. Flash column chromatography was performed with Merck silica gel (230–400 mesh). Reactions were carried out under air conditions.

2.2. Preparation of Fe3O4 Superparamagnetic Nanoparticle MNPs

Co-precipitation of FeCl3.6H2O and FeCl2·4H2O in ammonia was applied for MNP preparation, according to a procedure described elsewhere [39]. FeCl3·6H2O (5.41 g, 20 mmol) and FeCl2·4H2O (2 g, 10 mmol) were dissolved in 100 mL deionized water under argon atmosphere at 90 °C; then, ammonium hydroxide 25% solution (20 mL) was added with vigorous mechanical stirring. After 20 min, the color of the bulk solution turned black. The resulting black MNPs were separated by applying a permanent external magnet, washed three times with deionized water, and then dried under vacuum at 60 °C for 12 h.

2.3. Preparation of KF/Fe3O4

This catalyst was prepared by mixing KF (2.5 g) and Fe3O4 (0.5 g) in a mortar. The mixture was transferred to a flask and vacuum-dried in an oil bath at 120–130 °C for 7 h. The resulting KF/Fe3O4 was kept in a desiccator until required.

2.4. General Procedure for the Synthesis of C–N Materials

Copper (I) iodide (19 mg, 10 mol%) and KF/Fe3O4 (160 mg) were added to a magnetically stirred mixture of nitrogenated compound (1.5 mmol) and aryl halide (1 mmol) in DMF (5 mL). The reaction mixture was then heated at 120 °C for the specified time. The reaction progress was monitored by TLC. The reaction mixture was allowed to cool down to room temperature, filtered, and then partitioned between ethyl acetate (3 mL) and saturated aqueous NaCl solution (3 × 10 mL). The organic fraction was washed with water (3 × 10 mL), dried with sodium sulphate, filtered, and concentrated in a rotary evaporator. The crude product was purified by column chromatography using hexane and ethyl acetate (8:2) as eluent to afford the pure product. The melting points and spectroscopic data for the known products were identical to those reported in the literature.

2.5. General Procedure for the Synthesis of Diaryl Ethers

Procedure A. Cu-catalyzed coupling of aryl iodides with phenols in DMF solvent: A screwcap, oven-dried tube was charged with the corresponding phenol (1.3 mmol), iodobenzene (1 mmol), KF/Fe3O4 (160 mg), and CuI catalyst (0.1 mmol). The tube was septum-capped, air evacuated, and back-filled with argon. Anhydrous DMF (5 mL) was added, and the reaction mixture was stirred at 120 °C for 6 h (the progress of the reaction was monitored by TLC). After the complete disappearance of the iodobenzene, the mixture was cooled to room temperature, diluted with ethyl acetate, filtered, and the crude was purified by column chromatography on silica gel (ethyl acetate/hexane) to obtain the desired product.
Procedure B. Coupling of aryl fluorides/aryl chlorides with Phenols in DMF solvent: A mixture of the phenol (1.3 mmol), the appropriate aryl halide (1.2 mmol), and KF/Fe3O4 (160 mg), in DMF as a solvent (5 mL), was heated to 120 °C and stirred under air conditions until the completion of the reaction (monitoring by TLC or GC). The mixture was cooled to room temperature, diluted with ethyl acetate, and filtered. The resulting solution was placed in a separating funnel and washed twice with H2O. The organic layer was dried with anhydrous Na2SO4, filtered, and concentrated in vacuo. The resulting products were purified either by column chromatography on silica gel (ethyl acetate/hexane) or recrystallization (in methanol).

3. Results and Discussion

3.1. Characterization of the Catalyst

The X-ray diffraction pattern of magnetic nanoparticles (MNPs) can be seen in Figure 1. Sharp and strong peaks indicate good sample crystallinity. The patterns at 2Θ values 30.1°, 35.4°, 43.1°, 53.4°, 57°, and 62.6° can be assigned to (220), (311), (400), (422), (511) and (440) crystal planes in Fe3O4 cubic lattice, which agree with the standard Fe3O4 (Joint Committee on Powder Diffraction Standards Card No 19-0629).
Scanning Electron Microscopy (SEM) analysis (Figure 2) of the KF/Fe3O4 showed uniform-sized particles before and after four reaction cycles with spherical morphology with an average size range of 20 nm.

3.2. Formation of C–N Bond

In our study, we initially addressed the optimization of the ratio of KF to Fe3O4, taking as a model the synthesis of benzanilide by cross-coupling the reaction of iodobenzene and benzamide in the presence of CuI in DMF at reflux as solvent (Scheme 1). The results are shown in Table 1.
First, we used the CaO/Fe3O4 blend, but it resulted in low yields. So, we decided to change the base to KF. Since KF is moisture sensitive, we prepared dry KF/Fe3O4 under vacuum at 70 °C. Some research has been conducted in the presence of a CuI catalyst, a strong base or a ligand, but it is not similar to our method. The new system (KF/Fe3O4) produced promising results, justifying a more extended study on its scope.
As can be seen in Table 1, when increasing the impregnated proportion of KF to Fe3O4, the yield of product increases too; these results may be assigned from increased adsorption of fluoride anions on the Fe3O4 surface. In Entry 4, Table 1 contains the ratio of KF to Fe3O4 (5 g/1 g) selected for further experiments.
To gain insight into the effect that solvent, temperature, CuI load and L-proline have on the reaction outcome, several experiments were run using the model reaction shown in Scheme 1. It is worth noting that L-proline is known to act as a ligand in the catalytic system CuI/L-proline [40]. The conditions and results are shown in Table 2.
It is worth pointing out that the reaction between 1 and 2 mediated by either KF or Fe3O4 alone did not proceed. As the results in Table 2 show, it was found that the solvent has a remarkable influence on the reaction success. In either o-xylene, toluene, or THF, only a trace amount of product was obtained (Table 2, Entries 1–3), whereas, in dioxane, DMSO, and DMF, 20%, 80%, and 90% yield of the desired product was obtained, respectively (Table 2, Entries 4–6). Then, we performed this reaction in the presence of 10 mol% L-proline and 5 mol% and 10 mol% copper (I) iodide (Table 2, Entries 7 and 8), but no real improvement was found compared to the ligand-free experiment. Moreover, several experiments were carried out to check the influence of the temperature. At 80 °C and room temperature, low yield or traces of products were obtained (Table 2, Entries 11 and 12). Conducting the reaction at 120 °C with either 10 mol% or 15 mol%, excellent yields of CuI were obtained (Table 2, Entries 9 and 10); therefore, we chose the conditions shown in Entry 10 of Table 2.
After optimization of the protocol, we subjected aryl halides for C–N bond formation in the arylation of amides, carbazoles, and indoles. The reagents used are shown in Figure 3 and the results are listed in Table 3.
According to the results shown in Table 3, indole, carbazole, and amide compounds were successfully transformed to the corresponding N-aryl compounds. The reactions of carbazole and indole with iodo- and 4-nitroiodobenzenes under optimum reaction conditions gave excellent yields (Table 3, Entries 5–8). Moreover, the amidation reaction carried out under similar reaction conditions, gave good yields of N-phenylamide (Table 3, Entries 1–4).
In addition to the magnetic base, ligand-free protocols also have great efficiency in Ullmann coupling reactions. During the past few years, considerable research reported the efficiency of the Ullman reaction using copper salts with several ligands for the preparation of diaryl ethers [39,41,42,43,44,45,46,47,48,49,50,51,52].

3.3. Formation of C–O Bond

Next, we investigated the strategy by copper (I) salt as a catalyst and KF/Fe3O4 as a base/bed in the reaction of iodobenzene (1) with p-cresol (17) to form 1-methyl-4-phenoxybenzene (18) as a model procedure in different solvents (Scheme 2); the results are listed in Table 4.
As shown in Table 4, it was found that the solvent has a profound effect on the yield of the product. As before, the reaction was performed without a ligand. When we used o-xylene, toluene, and THF, only a trace amount of product was obtained (Table 4, Entries 1–3), but in dioxane, DMSO, and DMF, 30%, 80%, and 95% yields of the desired product were obtained, respectively (Table 4, Entries 4–6). First, to optimize the amount of catalyst, we examined 15 and 10 mol% of copper (I) iodide in refluxing DMF. The reaction was performed satisfactorily using these conditions and yielded corresponding N-aryl compounds (Table 4, Entries 6 and 10). Moreover, note that the use of 10 mol% of L-proline in conjunction with 10 mol% of CuI did not improve the reaction outcome (Table 4, compare Entries 12 and 10).
Using optimized reaction conditions (Table 4), we carried out the reaction between aryl iodides and phenols to extend the scope of this protocol of synthesis of diaryl ethers using the reagents shown in Figure 4. The results are listed in Table 5.
When the results from Entries 5 and 9 in Table 5, in which the same product (34) is formed, are compared, it becomes obvious that there is an accelerating effect when the electron-withdrawing group is in the iodide rather than in the phenol since the same yield is obtained in 4 h (Table 5, Entry 5) as opposed to 12 h (Table 5, Entry 9).
This protocol provides an outstanding entry to diaryl ethers, no matter whether the groups in either the iodide or the phenol are electron-withdrawing or electron-donating or neither. It is worth noting that when the phenolic compound carries an electron-withdrawing group, the reaction is slower but still proceeds with excellent yield (Table 5, Entry 9). In Entries 5, 6, and 8 in Table 5, 1-iodo-4-nitrobenzene reacted quantitatively in 4 h with phenols with no substitution or a p-alkyl moiety. Furthermore, when iodobenzene is reacted with phenol, p-cresol, 2-naphthol, and 4-tert-butylphenol, the reaction yields are excellent, taking 5 h to complete (Table 5, Entries 1–4); nevertheless, the reaction takes up to 12 h when the reactant is 4-nitrophenol, as mentioned before (Table 5, Entry 9). Finally, 1-iodo-4-methoxyphenol and p-cresol produce the corresponding diaryl ether in 90% yield in only 3.5 h (Table 5, Entry 7).
In continuation to our studies on the reaction between phenols and nitro-activated aryl halides, we investigated the scope of this reaction by coupling various phenolic nucleophiles with 1-fluoro- and 1-chloro-4-nitrobenzene. Remarkably, after optimizing the reaction conditions, we found that the reactions could not only proceed in excellent yields in air atmosphere but also with no need for a copper catalyst whatsoever. The base/bed, KF/Fe3O4, along with the right solvent choice, was enough to promote the reaction with these substrates (Table 6 and Table 7). It is worth noting that the hydrogen halide eventually formed will be neutralized and the excess molecules will be lost to the atmosphere, where they will react with moisture or oxygen.
As a result, it is shown that the nucleophilic aromatic substitution between active aryl halides and phenols can be performed under mild conditions to prepare diaryl ethers. The coupling of activated aryl fluorides with various phenols gives the corresponding product in excellent yields; when electron-rich groups are present in the para or orto position of phenols, excellent yields are obtained in a short time.

3.4. Role of the Copper and Recyclability of the Catalyst

As stated in the literature [33,38], the role of the ligand in this type of chemical transformation is to activate the copper catalyst to dissolve and start the reaction process. In the reaction that we have studied, the copper is activated in the presence of the solvent and enters the catalytic reaction cycle without the presence of a ligand. The reaction mechanism is shown in Figure 5. The reaction has a two-step catalytic cycle: oxidative addition and reductive elimination. In the first step, the copper catalyst is oxidized by bonding to the reactants, and in the second step, the reactants are converted to the products, and copper is reduced back and so the cycle continues.
The recyclability of the magnetic catalyst was examined for the reaction of iodobenzene (1a) and benzamide (2a) to give benzanilide (3a) and the synthesis of 1-(4-nitrophenoxy)-4-methylbenzene (3x) by a reaction between 1-chloro-4-nitrobenzene (1b) and p-cresol (2b). After the completion of either reaction, the catalyst could be simply separated with a magnet, and then recycled at least three times with no significant change in its catalytic activity (Table 8).

4. Conclusions

This work evaluated the subject of heteroatom coupling reactions with active alkyl halides in SNAr substitution reactions.
In conclusion, the copper-catalyzed coupling reaction can be performed using stable CuI as a copper source in the presence of KF/Fe3O4 as a base. This is a simple, inexpensive, and efficient method for the coupling of carbon and heteroatoms. We have developed a new and simple strategy for constructing carbon–heteroatom bonds, such as C–O and C–N bonds. The reaction is performed in the presence of KF/Fe3O4, which acts as a magnetic base, and CuI as a catalyst without using any ligand compared to the same synthetic procedure. This method provides an efficient route for the preparation of diaryl ethers from phenols and aryl halides. The formation of C–N bonds via the Ullmann protocol is also achieved. In fact, in this article, a new and effective method in coupling reactions is designed and optimized.
The advantage of this method over previous methods is that our reaction is performed in the presence of a nanomagnetic substrate (bed) that shows selective adsorption to potassium ions and other metal cations. Furthermore, the reaction was performed as an experimental technique in which no organic ligand was used, and the reaction was carried out under mild conditions. We conclude that the KF/Fe3O4 blend is able to produce a synthesis with high selectivity in coupling reactions. For example, compared to a catalyst such as KF/Al2O3, our material seems to be a better option for performing chemical reactions with the same reactive materials. We believe that KF supported on magnetic Fe3O4 provides an excellent complement to other bases in copper-catalyzed coupling reactions, which have already been utilized in several works.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/appliedchem1020007/s1. Figure S1: NMR spectra obtained for benzanilide (compound 3), Figure S2: NMR spectra obtained for N-Phenylindole (compound 13), Figure S3: NMR spectra obtained for N-Phenylcarbazole (compound 15), Figure S4: NMR spectrum obtained for 1-Methyl-4-phenoxybenzene (compound 18), Figure S5: NMR spectrum obtained for Diphenyl ether (compound 31), Figure S6: NMR spectra obtained for 1-Nitro-4-phenoxybenzene (compound 34), Figure S7: NMR spectra obtained for 4-Methyl 4-nitro-diphenyl ether (compound 35), Figure S8: NMR spectra obtained for 1-(4-Nitrophenoxy)-4-tert-butylbenzene (compound 37), Figure S9: NMR spectra obtained for 1-(2,4 Dichlorophenoxy)-4-nitrobenzene (compound 39), Figure S10: NMR spectra obtained for 1-(4-Chlorophenoxy)-4-nitrobenzene (compound 40), Figure S11: NMR spectra obtained for 1-(4-Nitrophenoxy)-4-nitrobenzene (compound 41), Figure S12: NMR spectrum obtained for 1-Bromo-4-(4-nitrophenoxy)benzene (compound 42).

Author Contributions

Conceptualization and methodology, M.T. and M.Q.; investigation, M.T., M.Q., A.R. and M.F.; writing—original draft preparation, M.T., M.Q., A.R., M.F., S.I., C.J.D.-V. and I.M.L.-C.; writing—review and editing, M.T., M.Q., S.I., C.J.D.-V. and I.M.L.-C.; supervision, M.T., M.Q., S.I., C.J.D.-V. and I.M.L.-C.; funding acquisition, C.J.D.-V. and I.M.L.-C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Regional Government “Junta de Extremadura” and the European Regional Development Fund (grants GR18035, GR18171, and IB16167); C.J.D.-V. and I.M.L.-C. acknowledge this funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article and Supplementary Material.

Conflicts of Interest

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

References

  1. Allen, S.E.; Walvoord, R.R.; Padilla-Salinas, R.; Kozlowski, M.C. Aerobic Copper-Catalyzed Organic Reactions. Chem. Rev. 2013, 113, 6234–6458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Evans, D.A.; Wood, M.R.; Trotter, B.W.; Richardson, T.I.; Barrow, J.C.; Katz, J.L. Total Syntheses of Vancomycin and Eremomycin Aglycons. Angew. Chem. Int. Ed. 1998, 37, 2700–2704. [Google Scholar] [CrossRef]
  3. Boger, D.L.; Miyazaki, S.; Kim, S.H.; Wu, J.H.; Castle, S.L.; Loiseleur, O.; Jin, Q. Total synthesis of the vancomycin aglycon. J. Am. Chem. Soc. 1999, 121, 10004–10011. [Google Scholar] [CrossRef]
  4. Procter, D.J. The synthesis of thiols, selenols, sulfides, selenides, sulfoxides, selenoxides, sulfones and selenones. J. Chem. Soc. Perkin Trans. 2001, 1, 335–354. [Google Scholar] [CrossRef]
  5. Ley, S.V.; Thomas, A.W. Modern Synthetic Methods for Copper-Mediated C(aryl)-O, C(aryl)-N, and C(aryl)-S Bond Formation. Angew. Chem.-Int. Ed. 2003, 42, 5400–5449. [Google Scholar] [CrossRef]
  6. Scott Sawyer, J. Recent advances in diaryl ether synthesis. Tetrahedron 2000, 56, 5045–5065. [Google Scholar] [CrossRef]
  7. Bariwal, J.; Van Der Eycken, E. C-N bond forming cross-coupling reactions: An overview. Chem. Soc. Rev. 2013, 42, 9283–9303. [Google Scholar] [CrossRef]
  8. Khalilzadeh, M.A.; Hosseini, A.; Pilevar, A. Potassium Fluoride Supported on Natural Nanoporous Zeolite: A New Solid Base for the Synthesis of Diaryl Ethers. Eur. J. Org. Chem. 2011, 2011, 1587–1592. [Google Scholar] [CrossRef]
  9. Agawane, S.M.; Nagarkar, J.M. Nano ceria catalyzed Ullmann type coupling reactions. Tetrahedron Lett. 2011, 52, 5220–5223. [Google Scholar] [CrossRef]
  10. Ono, Y.; Baba, T. Strong solid bases for organic reactions. In Catalysis; Royal Society of Chemistry: Cambridge, UK, 2000; Volume 15, pp. 1–39. ISBN 9781847553270. [Google Scholar]
  11. Albouy, D.; Laspéras, M.; Etemad-Moghadam, G.; Koenig, M. Role of base catalysts upon the Pudovik reaction: Unexpected synthesis of 1,2-dihydropyridine phosphonate derivatives. Tetrahedron Lett. 1999, 40, 2311–2314. [Google Scholar] [CrossRef]
  12. Smith, G.; Notheisz, F. Heterogeneous Catalysis in Organic Chemistry; Academic Press: San Diego, CA, USA, 1999; ISBN 9780126516456. [Google Scholar]
  13. Arzamendi, G.; Campo, I.; Arguiñarena, E.; Sánchez, M.; Montes, M.; Gandía, L.M. Synthesis of biodiesel with heterogeneous NaOH/alumina catalysts: Comparison with homogeneous NaOH. Chem. Eng. J. 2007, 134, 123–130. [Google Scholar] [CrossRef]
  14. Tedeschi, L.; Enders, D. Asymmetric synthesis of β-phosphono malonates via Fe2O3-mediated phospha-Michael addition to Knoevenagel acceptors. Org. Lett. 2001, 3, 3515–3517. [Google Scholar] [CrossRef]
  15. Chen, W.; Huang, Z.; Liu, Y.; He, Q. Preparation and characterization of a novel solid base catalyst hydroxyapatite loaded with strontium. Catal. Commun. 2008, 9, 516–521. [Google Scholar] [CrossRef]
  16. Motokura, K.; Fujita, N.; Mori, K.; Mizugaki, T.; Ebitani, K.; Kaneda, K. An acidic layered clay is combined with a basic layered clay for one-pot sequential reactions. J. Am. Chem. Soc. 2005, 127, 9674–9675. [Google Scholar] [CrossRef] [PubMed]
  17. Okachi, T.; Fujimoto, K.; Onaka, M. Practical carbonyl-ene reactions of α-methylstyrenes with paraformaldehyde promoted by a combined system of boron trifluoride and molecular sieves 4A. Org. Lett. 2002, 4, 1667–1669. [Google Scholar] [CrossRef] [PubMed]
  18. Perozo-Rondón, E.; Calvino-Casilda, V.; Martín-Aranda, R.M.; Casal, B.; Durán-Valle, C.J.; Rojas-Cervantes, M.L. Catalysis by basic carbons: Preparation of dihydropyridines. Appl. Surf. Sci. 2006, 252, 6080–6083. [Google Scholar] [CrossRef]
  19. Choi, M.K.W.; Yu, W.Y.; So, M.H.; Zhou, C.Y.; Deng, Q.H.; Che, C.M. A non-cross-linked soluble polystyrene-supported ruthenium catalyst for carbenoid transfer reactions. Chem.-An Asian J. 2008, 3, 1256–1265. [Google Scholar] [CrossRef] [PubMed]
  20. So, M.H.; Liu, Y.; Ho, C.M.; Che, C.M. Graphite-supported gold nanoparticles as efficient catalyst for aerobic oxidation of benzylic amines to imines and N-substituted 1,2,3,4-tetrahydroisoquinolines to amides: Synthetic applications and mechanistic study. Chem.-Asian J. 2009, 4, 1551–1561. [Google Scholar] [CrossRef] [PubMed]
  21. Lewis, L.N. Chemical Catalysis by Colloids and Clusters. Chem. Rev. 1993, 93, 2693–2730. [Google Scholar] [CrossRef]
  22. Trnka, T.M.; Grubbs, R.H. The development of L2X2RU=CHR olefin metathesis catalysts: An organometallic success story. Acc. Chem. Res. 2001, 34, 18–29. [Google Scholar] [CrossRef]
  23. Malinsky, M.D.; Kelly, K.L.; Schatz, G.C.; Van Duyne, R.P. Chain length dependence and sensing capabilities of the localized surface plasmon resonance of silver nanoparticles chemically modified with alkanethiol self-assembled monolayers. J. Am. Chem. Soc. 2001, 123, 1471–1482. [Google Scholar] [CrossRef]
  24. McConnell, W.P.; Novak, J.P.; Brousseau, L.C.; Fuierer, R.R.; Tenent, R.C.; Feldheim, D.L. Electronic and optical properties of chemically modified metal nanoparticles and molecularly bridged nanoparticle arrays. J. Phys. Chem. B 2000, 104, 8925–8930. [Google Scholar] [CrossRef]
  25. Shi, F.; Tse, M.K.; Pohl, M.M.; Radnik, J.; Brückner, A.; Zhang, S.; Beller, M. Nano-iron oxide-catalyzed selective oxidations of alcohols and olefins with hydrogen peroxide. J. Mol. Catal. A Chem. 2008, 292, 28–35. [Google Scholar] [CrossRef]
  26. Li, S.; Zhang, W.; So, M.H.; Che, C.M.; Wang, R.; Chen, R. One-pot solvothermal synthesis of Pd/Fe3O4 nanocomposite and its magnetically recyclable and efficient catalysis for Suzuki reactions. J. Mol. Catal. A Chem. 2012, 359, 81–87. [Google Scholar] [CrossRef]
  27. Beydoun, D.; Amal, R.; Low, G.K.C.; McEvoy, S. Novel Photocatalyst: Titania-Coated Magnetite. Activity and Photodissolution. J. Phys. Chem. B 2000, 104, 4387–4396. [Google Scholar] [CrossRef]
  28. Zhang, Y.-P.; Shi, A.-H.; Yang, Y.-S.; Li, C.-L. Impregnated copper on magnetite as catalyst for the O-arylation of phenols with aryl halides. Chin. Chem. Lett. 2014, 25, 141–145. [Google Scholar] [CrossRef]
  29. Hu, S.; Guan, Y.; Wang, Y.; Han, H. Nano-magnetic catalyst KF/CaO-Fe3O4 for biodiesel production. Appl. Energy 2011, 88, 2685–2690. [Google Scholar] [CrossRef]
  30. Nejati, K.; Ahmadi, S.; Nikpassand, M.; Kheirollahi Nezhad, P.D.; Vessally, E. Diaryl ethers synthesis: Nano-catalysts in carbon-oxygen cross-coupling reactions. RSC Adv. 2018, 8, 19125–19143. [Google Scholar] [CrossRef] [Green Version]
  31. Wen, M.; Qi, H.; Zhao, W.; Chen, J.; Li, L.; Wu, Q. Phase transfer catalysis: Synthesis of monodispersed FePt nanoparticles and its electrocatalytic activity. Colloids Surfaces A Physicochem. Eng. Asp. 2008, 312, 73–78. [Google Scholar] [CrossRef]
  32. Ames, L.L., Jr. The Cation Sieve Properties of Clinoptilolite. Am. Mineral. 1960, 45, 689–700. [Google Scholar]
  33. Hosseinzadeh, R.; Tajbakhsh, M.; Mohadjerani, M.; Alikarami, M. Copper-catalysed N-arylation of arylsulfonamides with aryl bromides and aryl iodides using KF/Al2O3. J. Chem. Sci. 2010, 122, 143–148. [Google Scholar] [CrossRef] [Green Version]
  34. Elazab, H.A.; Sadek, M.A.; El-Idreesy, T.T. Microwave-assisted synthesis of palladium nanoparticles supported on copper oxide in aqueous medium as an efficient catalyst for Suzuki cross-coupling reaction. Adsorpt. Sci. Technol. 2018, 36, 1352–1365. [Google Scholar] [CrossRef]
  35. Panahi, F.; Daneshgar, F.; Haghighi, F.; Khalafi-Nezhad, A. Immobilized Pd nanoparticles on silica-starch substrate (PNP-SSS): Efficient heterogeneous catalyst in Buchwald–Hartwig C–N cross coupling reaction. J. Organomet. Chem. 2017, 851, 210–217. [Google Scholar] [CrossRef]
  36. Hajipour, A.R.; Khorsandi, Z.; Fatemeh Mohammadi Metkazini, S. Palladium nanoparticles supported on cysteine-functionalized MNPs as robust recyclable catalysts for fast O- and N-arylation reactions in green media. J. Organomet. Chem. 2019, 899, 120793. [Google Scholar] [CrossRef]
  37. Veisi, H.; Safarimehr, P.; Hemmati, S. Buchwald–Hartwig C–N cross coupling reactions catalyzed by palladium nanoparticles immobilized on thio modified-multi walled carbon nanotubes as heterogeneous and recyclable nanocatalyst. Mater. Sci. Eng. C 2019, 96, 310–318. [Google Scholar] [CrossRef] [PubMed]
  38. Hosseinzadeh, R.; Tajbakhsh, M.; Alikarami, M. Copper-catalyzed N-arylation of diazoles with aryl bromides using KF/Al2O3: An improved protocol. Tetrahedron Lett. 2006, 47, 5203–5205. [Google Scholar] [CrossRef]
  39. Kim, K.D.; Kim, S.S.; Choa, Y.-H.; Kim, H.T. Formation and Surface Modification of Fe3O4 Nanoparticles by Co-precipitation and Sol-gel Method. J. Ind. Eng. Chem. 2007, 13, 1137–1141. [Google Scholar]
  40. Xie, X.; Cai, G.; Ma, D. CuI/L-proline-catalyzed coupling reactions of aryl halides with activated methylene compounds. Org. Lett. 2005, 7, 4693–4695. [Google Scholar] [CrossRef]
  41. Marcoux, J.F.; Doye, S.; Buchwald, S.L. A general copper-catalyzed synthesis of diaryl ethers. J. Am. Chem. Soc. 1997, 119, 10539–10540. [Google Scholar] [CrossRef]
  42. Cai, Q.; Zou, B.; Ma, D. Mild Ullmann-type biaryl ether formation reaction by combination of ortho-substituent and ligand effects. Angew. Chem.-Int. Ed. 2006, 45, 1276–1279. [Google Scholar] [CrossRef] [PubMed]
  43. Buck, E.; Song, Z.J.; Tschaen, D.; Dormer, P.G.; Volante, R.P.; Reider, P.J. Ullmann Diaryl Ether Synthesis: Rate Acceleration by 2,2,6,6-Tetramethylheptane-3,5-dione. Org. Lett. 2002, 4, 1623–1626. [Google Scholar] [CrossRef] [PubMed]
  44. Zhao, Y.; Wang, Y.; Sun, H.; Li, L.; Zhang, H. Ullmann reaction in tetraethyl orthosilicate: A novel synthesis of triarylamines and diaryl ethers. Chem. Commun. 2007, 3186–3188. [Google Scholar] [CrossRef]
  45. Miao, T.; Wang, L. Immobilization of copper in organic-inorganic hybrid materials: A highly efficient and reusable catalyst for the Ullmann diaryl etherification. Tetrahedron Lett. 2007, 48, 95–99. [Google Scholar] [CrossRef]
  46. Ma, D.; Cai, Q. N,N-Dimethyl Glycine-Promoted Ullmann Coupling Reaction of Phenols and Aryl Halides. Org. Lett. 2003, 5, 3799–3802. [Google Scholar] [CrossRef]
  47. Lv, X.; Bao, W. A β-keto ester as a novel, efficient, and versatile ligand for copper(I)-catalyzed C-N, C-O, and C-S coupling reactions. J. Org. Chem. 2007, 72, 3863–3867. [Google Scholar] [CrossRef] [PubMed]
  48. Bistri, O.; Correa, A.; Bolm, C. Iron-catalyzed C-O cross-couplings of phenols with aryl iodides. Angew. Chem.-Int. Ed. 2008, 47, 586–588. [Google Scholar] [CrossRef]
  49. Fagan, P.J.; Hauptman, E.; Shapiro, R.; Casalnuovo, A. Using intelligent/random library screening to design focused libraries for the optimization of homogeneous catalysts: Ullmann ether formation. J. Am. Chem. Soc. 2000, 122, 5043–5051. [Google Scholar] [CrossRef]
  50. Gujadhur, R.K.; Bates, C.G.; Venkataraman, D. Formation of aryl-nitrogen, aryl-oxygen, and aryl-carbon bonds using well-defined copper(I)-based catalysts. Org. Lett. 2001, 3, 4315–4317. [Google Scholar] [CrossRef] [PubMed]
  51. Naidu, A.B.; Jaseer, E.A.; Sekar, G. General, mild, and intermolecular Ullmann-type synthesis of diaryl and alkyl aryl ethers catalyzed by diol-copper(I) complex. J. Org. Chem. 2009, 74, 3675–3679. [Google Scholar] [CrossRef]
  52. Naidu, A.B.; Raghunath, O.R.; Prasad, D.J.C.; Sekar, G. An efficient BINAM-copper(II) catalyzed Ullmann-type synthesis of diaryl ethers. Tetrahedron Lett. 2008, 49, 1057–1061. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the MNPs.
Figure 1. XRD patterns of the MNPs.
Appliedchem 01 00007 g001
Figure 2. SEM Images of synthesized (a) KF/Fe3O4 and (b) KF/Fe3O4 after 4 cycles.
Figure 2. SEM Images of synthesized (a) KF/Fe3O4 and (b) KF/Fe3O4 after 4 cycles.
Appliedchem 01 00007 g002
Scheme 1. Model reaction for the C–N bond formation.
Scheme 1. Model reaction for the C–N bond formation.
Appliedchem 01 00007 sch001
Figure 3. Reagents used for arylation of amides.
Figure 3. Reagents used for arylation of amides.
Appliedchem 01 00007 g003
Scheme 2. Synthesis of diaryl ethers.
Scheme 2. Synthesis of diaryl ethers.
Appliedchem 01 00007 sch002
Figure 4. Reagents used for the synthesis of diaryl ethers.
Figure 4. Reagents used for the synthesis of diaryl ethers.
Appliedchem 01 00007 g004
Figure 5. Catalytic cycle.
Figure 5. Catalytic cycle.
Appliedchem 01 00007 g005
Table 1. Optimization of the KF to Fe3O4 ratio in the synthesis of benzanilide (3) [a].
Table 1. Optimization of the KF to Fe3O4 ratio in the synthesis of benzanilide (3) [a].
EntryKF (g)Fe3O4 (g)KF/Fe3O4 (g:g)Yield % [b]
10.510.520
210.5260
320.5485
42.50.5590
530.5690
[a] Reaction Conditions: PhI (1) (1 mmol), Benzamide (2) (1.5 mmol), refluxing DMF (5 mL), CuI (15 mol%). [b] Isolated yield.
Table 2. Optimization of the reaction conditions for the synthesis of benzanilide (3) [a].
Table 2. Optimization of the reaction conditions for the synthesis of benzanilide (3) [a].
EntrySolventT (°C)CuI (mol%)L-Proline (mol%)Yield (%) [b]
1o-XyleneReflux15-Trace
2TolueneReflux15-Trace
3THFReflux15-Trace
4DioxaneReflux15-20
5DMSOReflux15-80
6DMFReflux15-90
7DMFReflux101090
8DMFReflux51060
9DMF12015-90
10DMF12010-90
11DMF8015-30
12DMFr.t.15-Trace
[a] Reaction conditions: Iodobenzene (1) (1 mmol), benzamide (2) (1.5 mmol), solvent (5 mL), KF/Fe3O4 (160 mg). [b] Isolated yield.
Table 3. CuI-catalyzed coupling of aryl halide and nitrogen containing compounds [a].
Table 3. CuI-catalyzed coupling of aryl halide and nitrogen containing compounds [a].
EntryAryl IodideNitrogenated CompoundProductTime (h)Yield (%) [b]
112 Appliedchem 01 00007 i001
3
590
215 Appliedchem 01 00007 i002
10
790
316 Appliedchem 01 00007 i003
11
780
417 Appliedchem 01 00007 i004
12
685
518 Appliedchem 01 00007 i005
13
790
648 Appliedchem 01 00007 i006
14
690
719 Appliedchem 01 00007 i007
15
390
849 Appliedchem 01 00007 i008
16
290
[a] Reaction conditions: Aryl iodide (1 mmol), nitrogenated compound (1.5 mmol), DMF (5 mL), 120 °C, CuI (10 mol%), KF/Fe3O4 (160 mg). [b] Isolated yield.
Table 4. Optimization of the reaction conditions for the synthesis of 1-methyl-4-phenoxybenzene (18) [a].
Table 4. Optimization of the reaction conditions for the synthesis of 1-methyl-4-phenoxybenzene (18) [a].
EntrySolventT (°C)CuI (mol%)L-Proline (mol%)Yield (%) [b]
1o-XyleneReflux15-Trace
2TolueneReflux15-Trace
3THFReflux15-Trace
4DioxaneReflux15-30
5DMSOReflux15-80
6DMFReflux15-95
7DMF12015-95
8DMF8015-30
9DMFr.t.15-Trace
10DMF12010-95
11DMF1205570
12DMF120101095
[a] Reaction conditions: PhI (1) (1 mmol), p-Cresol (17) (1.3 mmol), solvent (5 mL), KF/Fe3O4 (160 mg), 3 h in argon atmosphere. [b] Isolated yield.
Table 5. Coupling of aryl iodides with phenols catalyzed by CuI [a].
Table 5. Coupling of aryl iodides with phenols catalyzed by CuI [a].
EntryAryl IodidePhenolProductTime (h)Yield (%) [b]
1122 Appliedchem 01 00007 i009
31
595
2117 Appliedchem 01 00007 i010
18
595
3123 Appliedchem 01 00007 i011
32
588
4124 Appliedchem 01 00007 i012
33
5.595
5425 Appliedchem 01 00007 i013
34
495
6417 Appliedchem 01 00007 i014
35
492
71817 Appliedchem 01 00007 i015
36
3.590
8424 Appliedchem 01 00007 i016
37
495
9125 Appliedchem 01 00007 i017
34
1295
[a] Reaction conditions: Aryl iodide (1 mmol), phenol (1.3 mmol), DMF (5 mL), 120 °C, CuI (20 mg), KF/Fe3O4 (160 mg), in argon atmosphere. [b] Isolated yield.
Table 6. Coupling of 1-fluoro-4-nitrobenzene (20) with phenols catalyzed by KF/Fe3O4 [a].
Table 6. Coupling of 1-fluoro-4-nitrobenzene (20) with phenols catalyzed by KF/Fe3O4 [a].
EntryPhenolProductTime (min)Yield (%) [b]
122 Appliedchem 01 00007 i018
34
1095
217 Appliedchem 01 00007 i019
35
1095
323 Appliedchem 01 00007 i020
38
2588
424 Appliedchem 01 00007 i021
37
5095
526 Appliedchem 01 00007 i022
39
3095
627 Appliedchem 01 00007 i023
40
2092
725 Appliedchem 01 00007 i024
41
6090
828 Appliedchem 01 00007 i025
42
2595
929 Appliedchem 01 00007 i026
43
1595
1030 Appliedchem 01 00007 i027
44
4595
[a] Reaction conditions: Aryl fluoride (1.2 mmol), phenol (1.3 mmol), DMF (5 mL), 120 °C, KF/Fe3O4 (160 mg). [b] Isolated yield.
Table 7. Coupling of 1-chloro-4-nitrobenzene (21) with phenols catalyzed by KF/Fe3O4 [a].
Table 7. Coupling of 1-chloro-4-nitrobenzene (21) with phenols catalyzed by KF/Fe3O4 [a].
EntryPhenolProductTime (h)Yield (%) [b]
122341.590
217350.995
323384.585
42437590
52639395
627404.590
72541585
828424.592
929431.594
103044390
[a] Reaction conditions: Aryl chloride (1.2 mmol), phenol (1.3 mmol), DMF (5 mL), 120 °C, KF/Fe3O4 (160 mg). [b] Isolated yield.
Table 8. Recyclability of KF/Fe3O4 in the synthesis of 3 [a] and 35 [b].
Table 8. Recyclability of KF/Fe3O4 in the synthesis of 3 [a] and 35 [b].
Reaction CycleYield (%) 3 [c]Yield (%) 35 [c]
19095
29095
38890
48488
[a] Reaction conditions: Iodobenzene (1) (1 mmol), benzamide (2) (1.5 mmol), DMF (5 mL), 120 °C, CuI 10 mol%, KF/Fe3O4 (160 mg). [b] Reaction conditions: 1-Iodo-4-nitrobenzene (4) (1.2 mmol), p-cresol (17) (1.3 mmol), DMF (5 mL), 120 °C, KF/Fe3O4 (160 mg). [c] Isolated yield.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tajbakhsh, M.; Ramezani, A.; Qandalee, M.; Falahati, M.; Durán-Valle, C.J.; Izquierdo, S.; López-Coca, I.M. Carbon–Heteroatom Bond Formation via Coupling Reactions Performed on a Magnetic Nanoparticle Bed. AppliedChem 2021, 1, 75-89. https://doi.org/10.3390/appliedchem1020007

AMA Style

Tajbakhsh M, Ramezani A, Qandalee M, Falahati M, Durán-Valle CJ, Izquierdo S, López-Coca IM. Carbon–Heteroatom Bond Formation via Coupling Reactions Performed on a Magnetic Nanoparticle Bed. AppliedChem. 2021; 1(2):75-89. https://doi.org/10.3390/appliedchem1020007

Chicago/Turabian Style

Tajbakhsh, Mahmood, Ali Ramezani, Mohammad Qandalee, Mobina Falahati, Carlos J. Durán-Valle, Silvia Izquierdo, and Ignacio M. López-Coca. 2021. "Carbon–Heteroatom Bond Formation via Coupling Reactions Performed on a Magnetic Nanoparticle Bed" AppliedChem 1, no. 2: 75-89. https://doi.org/10.3390/appliedchem1020007

Article Metrics

Back to TopTop