Next Article in Journal
In-Depth Study of Laser Diode Ablation of Kapton Polyimide for Flexible Conductive Substrates
Next Article in Special Issue
Effects of Angular Dependency of Particulate Light Scattering Intensity on Determination of Samples with Bimodal Size Distributions Using Dynamic Light Scattering Methods
Previous Article in Journal
Synthesis and Surface-Enhanced Raman Scattering of Ultrathin SnSe2 Nanoflakes by Chemical Vapor Deposition
Previous Article in Special Issue
Bioinspired Design of Alcohol Dehydrogenase@nano TiO2 Microreactors for Sustainable Cycling of NAD+/NADH Coenzyme
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Iodoxybenzoic Acid Supported on Multi Walled Carbon Nanotubes as Biomimetic Environmental Friendly Oxidative Systems for the Oxidation of Alcohols to Aldehydes

by
Bruno Mattia Bizzarri
1,
Issam Abdalghani
2,
Lorenzo Botta
1,
Anna Rita Taddei
3,
Stefano Nisi
4,
Marco Ferrante
4,5,
Maurizio Passacantando
2,
Marcello Crucianelli
2,* and
Raffaele Saladino
1,*
1
Department of Ecology and Biology, University of Tuscia, Largo dell’Università, 01100 Viterbo (VT), Italy
2
Department of Physical and Chemical Sciences, University of L’Aquila, Via Vetoio, I-67100 Coppito (AQ), Italy
3
Great Equipment Center Electron Microscopy Section, University of Tuscia, Largo dell’Università, 01100 Viterbo (VT), Italy
4
Laboratori Nazionali del Gran Sasso, Via G. Acitelli, 22, 67100 Assergi (AQ), Italy
5
Trace Research Centre, Via I. Silone 6, 64015 Nereto (TE), Italy
*
Authors to whom correspondence should be addressed.
Nanomaterials 2018, 8(7), 516; https://doi.org/10.3390/nano8070516
Submission received: 13 June 2018 / Revised: 5 July 2018 / Accepted: 6 July 2018 / Published: 10 July 2018
(This article belongs to the Special Issue Nanomaterials in Biocatalyst)

Abstract

:
Iodoxybenzoic acid (IBX) supported multi walled carbon nanotube (MWCNT) derivatives have been prepared as easily recyclable solid reagents. These compounds have been shown to be able to mimic the alcohol dehydrogenases and monooxygenases promoted oxidation of aromatic alcohols to corresponding aldehydes. Their reactivity was found to be dependent on the degree of functionalization of MWCNTs as well as from the chemical properties of the spacers used to bind IBX on the surface of the support. Au-decorated MWCNTs and the presence of longer spacers resulted in the optimal experimental conditions. A high conversion of the substrates and yield of desired products were obtained.

Graphical Abstract

1. Introduction

The oxidation of alcohols to corresponding carbonyl compounds is one of the most fundamental and important processes in synthetic organic chemistry. Although a variety of methods and reagents have been developed, they all suffer from the difficulty of selectively oxidizing primary alcohols to aldehydes without the concomitant formation of carboxylic acids and other over-oxidation products [1]. The oxidation of alcohols to aldehydes is usually performed in the presence of stoichiometric reagents [2] including the Dess–Martin oxidation [3], the Swern and Corey-Kim reaction [4], and the Burgess reagent [5]. Heavy metal reagents have been also used in catalytic procedures, for instance, hydrogen-transfer reactions (Ru, Rh, Ir) [6], and Oppenauer oxidations (Al, Zr, lanthanides) [4]. On the other hand, metal-free oxidations are desired processes in the context of green-chemistry due to the known toxicity and high environmental impact of metal species. In this context, biotechnological applications of oxidative enzymes, e.g., alcohol dehydrogenases and monooxygenases with high environmental compatibility, have been widely evaluated [7]. Unfortunately, there are drawbacks related to the use of these enzymes in large scale applications, encompassing the necessity to regenerate Nicotinamide Adenine Dinucleotide NAD(P)+ for alcohol dehydrogenases, and the use of low molecular weight redox mediators (such as 2,2,6,6-tetramethylpiperidin-1-yl-oxidanyl (TEMPO) and 1-hydroxybenzotriazole (HOBt)) in the case of monooxygenases [8]. For example, the “coupled substrate approach”, proposed for the regeneration of NAD(P)+ through the use of ketones and aldehydes as co-substrates, has shown close equilibrium conditions (that is low conversion of the substrate), complex purification procedures, and enzyme inhibition [9]. Alternative procedures, namely the “coupled enzyme approach”, where the NAD(P)+ regeneration is performed by a second enzyme, are disadvantaged by the complex realization of two-enzyme kinetic processes and by the exclusive enzyme specificities for the substrate and co-substrate [10]. Thus, continuous interest is devoted to developing biomimetic and environmentally friendly metal-free solid supported reagents for the selective oxidation of alcohols that are able to mimic a biological material in its structure or function [11].
In the last few years, solid-supported ortho-iodoxybenzoic acid (1-hydroxy-1λ5,2-benziodoxol-1,3-dione, IBX) reagents able to convert primary alcohols to corresponding aldehydes, have been prepared by the immobilization of the active iodine species on chemically inert supports including silica [12], polystyrene beads [13], and ionic liquids [14]. These reagents have been reported to be biomimetic [15,16], contemporary solving problems associated with insolubility in common organic solvents and low stability towards moisture of IBX, combining the environmental advantages of simple recyclability of the reagent, simple purification procedures, easier reaction optimization, and safety related issues [13,17]. For these reasons, IBX supported reagents have been recognized as environmental friendly reagents [18,19]. Single layer two-dimensional sp2 carbons graphene oxide (GO) has been also successfully applied as a solid support for the immobilization of IBX [20].
Multi Walled Carbon Nanotubes (MWCNTs) consist of several layers of graphene sheets rolled up into a cylindrical shape surrounding a central tube with lengths in the micrometer scale and diameters up to 100 nm [21]. They offer a high surface area for loading of the active species as well as biocompatibility and mechanical resistance [20]. Notably, the reactivity of supported active species can be tuned, moving from GO to MWCNTs as a consequence of the local curvature and of the specific physical and chemical properties of the support [22]. Here, we describe the preparation of IBX supported MWCNT solid reagents as biomimetic and environmentally friendly oxidative systems. Two different types of supports were investigated, MWCNTs and, an alternative, Au-decorated MWCNTs. The direct formation of amide (spacer-mediated) linkages with IBX and the high binding affinity of sulfur containing IBX derivatives for the Au surface were both applied for the immobilization of the iodine active reagent. The novel IBX supported MWCNT systems showed high reactivity and selectivity in the oxidation of primary alcohol to corresponding aldehydes, the efficacy of the system being controlled by the nature of the support and the length of the spacer.

2. Materials and Methods

2.1. Materials

Alcohols 18, organic solvents, HAuCl4, NaBH4, and GH Polypro membrane filters were purchased from Sigma-Aldrich (Saint Louis, MO, USA) and used without further purification. Gas chromatography mass spectrometry (GC–MS) was recorded on a Varian 410 GC-320 MS (Palo Alto, CA, USA) using a VF-5 ms column (30 m, 0.25 mm, 0.25 µm), and an electron beam of 70 eV. All experiments were done in triplicate. Ultrapure HNO3 and HCl obtained from a sub-boiling system (DuoPUR, Milestone, Bergamo, Italy) and ultrapure 18.2 MΩ water from a Milli-Q (Millipore, Burlington, MA, USA) were used for the sample dissolution. X-ray photoelectron spectroscopy (XPS) and inductively coupled plasma mass spectrometry (ICP–MS) were performed through an ultrahigh vacuum PHI 1257 system and an Agilent 7500 ICP–MS instrument under clean room ISO6 (Santa Clara, CA, USA), respectively.

2.2. Preparation of oxMWCNTs I

In a round-bottomed flask, equipped with an egg-shaped magnetic stirring bar, MWCNTs and a mixture of concentrated H2SO4–HNO3 (3:1) were stirred for 4.0 h at r.t. and an additional 12 h at 40 °C. The reaction mixture was cooled down to r.t. and cold H2O (400 mL) was poured into the reactor. The mixture was washed by centrifugation at 4000× g rpm (30 min), and the supernatant was removed. The remaining solid was further washed with deionized H2O (200 mL). At each washing step, the mixture was centrifuged (4000 rpm for 30 min), filtered using GH Polypro membrane filters 0.2 μm and the supernatant was removed. The resulting oxidized MWCNTs (oxMWCNTs I) were dried in vacuo and used without further purification.

2.3. Preparation of Oxidizing Solid Reagents IV A–B

oxMWCNTs I were suspended in Dimethyl Formamide DMF (0.8 mg/mL) and treated with N,N-diisopropylcarbodiimide (DIC; 250 mg, 2 mmol), HOBt (270 mg, 2 mmol), and N,N-diisopropyl ethylamine (DIPEA; 700 μL, 4 mmol) in a round-bottomed flask with an egg-shaped magnetic stirring bar for 15 min. Thereafter, the appropriate diamine (1,2-di-aminoethane for IIA and 1,6-diaminoethane for IIB) (2.0 mmol) was added to the solution under stirring for 12 h at 30 °C. The resulting solution was washed with DMF (5 × 5.0 mL) by centrifugation (4000× g rpm, 20 min) and filtered using GH Polypro membrane filters 0.2 μm. The resulting NH2–MWCNTs II AB were dried in vacuo. Successively, NH2–MWCNTs II AB (200 mg) suspended in DMF (0.8 mg/mL) were treated with DIC (790 mg, 6 mmol), and DIPEA (2.1 mL, 12 mmol), in a 500 mL round-bottomed flask with an egg-shaped magnetic stirring bar. Thereafter, 2-Iodo Benzoic Acid IBA (1.5 g, 6.0 mmol) was added to the solution and the mixture stirred for 8 h at 30 °C. The resulting IBA–MWCNTs III AB were washed with DMF and H2O by centrifugation (4000× g rpm, 20 min) and filtered using GH Polypro membrane filters 0.2 μm. IBA–MWCNTs III AB were then suspended in H2O (125 mg/250 mL) in a round-bottomed flask, and added to Oxone® (950 mg, 1.5 mmol) and methane sulfonic acid (100 μL, 1.5 mmol) under stirring for 8 h at r.t. Thereafter, IBX–MWCNTs IV AB were washed with DMF (5 × 5.0 mL) and H2O (3 × 5.0 mL) and filtered using GH Polypro membrane filters 0.2 μm.

2.4. Preparation of Oxidizing Solid Reagents VIII A–B

MWCNTs (100 mg) were sonicated in 100 mL of ethanol for 2 h. Afterwards, 8.5 mL of 0.1 M HAuCl4 ethanolic solution was added. In order to obtain Au particles, reduction with 300 mg of NaBH4 was carried out by stirring for about 30 min. Then, Au–MWCNTs V was isolated by centrifugation and filtered using GH Polypro membrane filters 0.2 μm washed several times with ethanol and dried at 80 °C. 2-amino-1-ethanethiol (for NH2–Au–MWCNTs VI A) and 6-amino-1-hexanthiol (for NH2-Au-MWCNTs VI B) was dissolved in a mixture of water (20 mL) and 1.0 M HCl (3.0 mL). Au–MWCNTs V (30 mg) and ethanol (3.0 mL) were added and the mixture was left under magnetic stirring for 24 h. After that time, the product was isolated by centrifugation, washed three times with 0.01 M NaOH and ethanol, and filtered using GH Polypro membrane filters 0.2 mm. Resulting NH2–Au–MWCNTs VI AB were dried under argon stream. NH2–Au–MWCNTs VI AB (200 mg) were suspended in DMF (0.8 mg/mL) and treated with DIC (790 mg, 6 mmol) and DIPEA (2.1 mL, 12 mmol) in a 500 mL round-bottomed flask with an egg-shaped magnetic stirring bar. Thereafter, IBA (1.5 g, 6 mmol) was added to the solution, and the mixture was stirred for 8 h at 30 °C. The resulting IBA–Au–MWCNTs VII AB were washed with DMF and H2O by centrifugation (4000× g rpm, 20 min) and filtered using GH Polypro membrane filters 0.2 μm. IBA–Au–MWCNTs VII AB were suspended in H2O (125 mg/250 mL) in a round-bottomed flask, then Oxone® (950 mg, 1.5 mmol), and methane sulfonic acid (100 μL, 1.5 mmol) were added and stirred for 8 h at r.t. Thereafter, IBX–Au–MWCNTs VIII AB were washed with DMF (5 × 10 mL) and H2O (3 × 10 mL) and filtered using GH Polypro membrane filters 0.2 μm.

2.5. Preparation of Oxidizing Solid Reagent VIII–C

11-mercapto-1-undecanol was dissolved in a mixture of water (20 mL) and 1.0 M HCl (3.0 mL). Au–MWCNTs V (30 mg) and ethanol (3.0 mL) were added and the mixture was left under magnetic stirring for 24 h. After that time, the product was isolated by centrifugation, washed three times with 0.01 M NaOH and ethanol, and filtered using GH Polypro membrane filters 0.2 mm. The resulting OH–Au–MWCNTs VI C was dried under argon stream. OH–Au–MWCNTs VI C (200 mg) was suspended in DMF (0.8 mg/mL) and treated with DIC (790 mg, 6 mmol), and DIPEA (2.1 mL, 12 mmol) in a 500 mL round-bottomed flask with an egg-shaped magnetic stirring bar. Thereafter, IBA (1.5 g, 6.0 mmol) was added to the solution and the mixture stirred for 8 h at 30 °C. The resulting IBA–Au–MWCNTs VII C was washed with DMF and H2O by centrifugation (4000× g rpm, 20 min) and filtered using GH Polypro membrane filters 0.2 μm. IBA–Au–MWCNTs VII C was suspended in H2O (125 mg/250 mL) in a round-bottomed flask, then Oxone® (950 mg, 1.5 mmol) and methane sulfonic acid (100 μL, 1.5 mmol) were added and stirred for 8 h at r.t. Thereafter, IBX–Au–MWCNTs VIII C was washed with DMF (5 × 10 mL) and H2O (3 × 10 mL) and filtered using GH Polypro membrane filters 0.2 μm.

2.6. Transmission Electron Microscopy (TEM), Scanning Electron Microscopy (SEM), and X-Ray Photoelectron Spectroscopy (XPS) Analyses

For transmission electron microscopy (TEM), samples were suspended in bi-distilled water. Droplets of sample suspensions (10 µL) were placed on formvar–carbon coated grids and allowed to adsorb for 60 s. Excess liquid was removed gently by touching the filter paper. Samples were observed with a JEOL 1200 EX II electron microscope (Waltham, MA, USA). Micrographs were acquired by the Olympus SIS VELETA CCD camera equipped with iTEM software (Waltham, MA, USA). For scanning electron microscopy (SEM), the sample suspensions (50 µL) were let to adsorb onto carbon tape attached to aluminum stubs and air dried at 25 °C. The observation was made by a JEOL JSM 6010LA electron microscope (Waltham, MA, USA) using Scanning Electron (SE) and Back Scattered Electrons (BSE) detectors. Energy Dispersive Spectroscopy (EDS) analysis was carried out to reveal the chemical elements. X-ray photoelectron spectroscopy (XPS) analysis was done in an ultrahigh vacuum PHI 1257 system equipped with a hemispherical analyzer, operating in the constant pass energy mode (with the total energy resolution of 0.8 eV) and using a non-monochromatized Mg Kα radiation source. The distance between the sample and the anode was about 40 mm, the illumination area was about 1 × 1 cm2, and the analyzed area was 0.8 × 2.0 mm2 with a take-off angle between the sample surface and the photoelectron energy analyzer of 45°. The energy scale was calibrated with reference to the binding energy of the C 1s at 284.8 eV with respect to the Fermi level. Survey scans of the III–B, IV–B, VII–A, and VIII–A compounds acquired in the range of 0–1100 eV (not shown here) displayed the contribution coming from the main elements involved in the reaction process for all of the samples: carbon, nitrogen, oxygen, sulfur, gold, and iodine. No contaminant species were observed within the sensitivity of the technique.

2.7. Inductively Coupled Plasma Mass-Spectrometry (ICP–MS) Analysis

The samples were weighed (from 1.6 to 6.9 mg) and transferred in Fluorinated ethylene propylene (FEP) vials, previously washed to avoid any kind of external contamination. Regia solution was chosen for the mineralization as it combines the oxidizing capacities of HNO3 with the complexing capacities of chlorides against I2 produced during digestion. In particular, 750 µL of HCl and 150 µL HNO3 were added and the solution was heated to 80 °C for 3 hours. The volume was adjusted to 5.0 mL and then diluted another 10 times before the ICP–MS analysis. The analysis was performed with an Agilent 7500 ICP–MS instrument (Palo Alto, CA, USA). Four standards at 10, 20, 50, and 100 ppb of iodine and gold were used for calibrating the instrument.

2.8. Oxidation of Aromatic Alcohols

The oxidation of alcohols 18 (1.0 mmol) in EtOAc (10 mL) was performed by adding the appropriate solid reagent (IV AB or VIII AC, 1.2 eq) to a single neck round-bottomed flask equipped with a water condenser under magnetic stirring at reflux conditions (c.a. 80° C) for 24 h. At the end of the oxidation, IV AB and VIII AC were filtered off using GH Polypro membrane filters 0.2 μm and washed with EtOAc (5 × 10 mL). The yield of aldehydes 916 was determined by GC–MS analysis using n-dodecane (0.1 mmol) as an internal standard. The reactions were performed in triplicate. GC–MS was performed using a VF-5ms column (30 m, 0.25 mm, 0.25 µm) through the following program: injection temperature 280 °C, detector temperature 280 °C, gradient 50 °C for 2 min, and 10 °C/min for 60 min, flow velocity of the carrier (helium), 1.0 mL min−1. In order to identify the structures of the products, two strategies were followed. First, the spectra of identifiable peaks were compared with commercially available electron mass spectrum libraries such as that of National Institute of Standards and Technology (NIST-Fison, Manchester, UK). In this latter case, spectra with at least 98% similarity were chosen. Secondly, GC–MS analysis was repeated using commercially available standard compounds. The original mass spectra of compounds 916 are reported in Figure S1 (Supporting Information).

3. Results

3.1. Preparation of IBX Supported MWCNTs and MWCNTs–Au Oxidizing Solid Reagents

The immobilization of IBX on MWCNTs was first based on the formation of an amide-type linkage between the spacer functionalized MWCNTs and 2-iodobenzoic acid (IBA), followed by activation of IBA to IBX (Scheme 1). In particular, commercially available MWCNTs were oxidized with HNO3/H2SO4 to oxMWCNTs I with the aim of increasing the amount of polar moieties (alcoholic and acidic groups) on the surface [23]. Next, ox-MWCNTs I was functionalized with selected alkyl diamino spacers (1,2-di-aminoethane and 1,6-diaminoethane) by coupling with N,N-diisopropyl carbodiimide (DIC) and 1-hydroxy benzotriazole (HOBt) in DMF at room temperature for 24 hours to yield the intermediates II AB. The effectiveness of the coupling procedure was confirmed by Fourier Transform Infrared Spectroscopy (FTIR) analysis for II-A as a selected example. In particular, the peak at 1649 cm−1, corresponding to the stretching vibration of the carboxylic groups in oxMWCNTs I (Figure S2), was shifted to 1633 cm−1 in II-A as a consequence of the amide formation, in accordance with data previously reported for the functionalization of MWCNTs (Figure S3) [24]. The intermediates II AB were successively suspended in DMF and treated with IBA at room temperature for 24 hours in the presence of DIC and HOBt to afford IBA–MWCNTs III AB. The formation of the novel amide linkage was again confirmed by the shift of the amide peak from 1633 cm−1 to 1627 cm−1 (Figure S4). Finally, III AB were activated to IBX-MWCNTs IV AB by reaction with Oxone® and methansulfonic acid. In this latter case, only a slight shift of the amide peak toward 1606 cm−1 was observed (Figure S5) [20].
As an alternative, Au decorated Au–MWCNTs V were used instead of oxMWCNTs I as anchorage supports. Briefly, Au–MWCNTs V [25] were treated with selected alkyl mercapto-amino spacers (2-amino-1-ethanethiol and 6-amino-1-hexanthiol, respectively) in an acidic water/ethanol mixture (pH 2, HCl 1.0 M) to afford the intermediates NH2–Au–MWCNTs VI AB by formation of covalent Au–sulfur bonds (Scheme 2). These intermediates were successively suspended in DMF and treated with IBA at room temperature for 24 h in the presence of DIC and HOBt to yield IBA–Au–MWCNTs VII AB. Finally, IBX–Au–MWCNTs VIII AB were obtained through the reaction of VII AB with Oxone® and methansulfonic acid [20]. The TEM images of IV B and VIII B, as the selected samples, are reported in Figure 1 (Panel A and C). In VIII B, the black-spots represent the Au particles, whose presence was unambiguously confirmed by SEM associated to BSE analysis (Figure S6). Note that the structural integrity of the MWCNTs was retained after the loading of IBX.
Moreover, VIII C was prepared using a longer thio-alcohol spacer (11-mercapto-1-undecanol), with the aim to bind IBA through the formation of an ester bond instead of an amide bond (Scheme 3). Briefly, Au–MWCNTs V was treated with 11-mercapto-1-undecanol in HCl 1.0 M and EtOH to afford the intermediates VI C by formation of covalent Au–sulfur bonds (Scheme 3). This intermediate was successively treated with DIC, DIPEA, and IBA to yield VII C. Finally, VII C was suspended in H2O and treated with Oxone® and methansulfonic acid to afford VIII C.

3.2. Determination of the Electron Binding Energies of the Elements by XPS Analysis

Figure 2 presents the detailed spectra of the C 1s, O 1s, N 1s, S 2p, Au 4f, and I 3d peaks of III B, IV B, VII A, and VIII A. All spectra were normalized to C 1s, which corresponded to the signal due to the MWCNTs support. In this way, we have the possibility of comparing the different peaks. XPS analysis clearly confirmed the presence of iodine and gold in the analyzed samples. Therefore, from the intensity of the XPS peaks (Figure 2) after the last step of the sample preparation (III BIV B and VII AVIII A), a slight leaching of Au and I was observed.
The C 1s spectra were fitted by the sum of five components assigned to C atoms belonging to: aromatic rings carbon (C=C/C–C, 284.8 eV), hydroxyl groups (C–OH, 285.9 eV), epoxy groups (C–O–C, 286.9 eV), carbonyl groups (C=O, 288.2 eV), and carboxyl groups (C=O(OH), 289.3 eV) (the hump at 290.6 eV was assigned to a π–π* shake-up satellite (in line with [20]). The O 1s spectra were fitted by the sum of three components: OH–C (533.4 eV), C–O–C (532 eV), and O=C (530.4 eV) [26]. Electron binding energies of the peak positions of N 1s, S 2p3/2, Au 4f7/2, and I 3d5/2 for all samples are listed in Table 1.

3.3. Determination of the Iodine Loading Factor by ICP–MS Analysis

The iodine Loading Factor (LF) for IV AB and VIII AC, defined as mmol of iodine per gram of support, was measured by Inductively Coupled Plasma Mass-Spectrometry (ICP–MS) analysis (Table 2). As reported in Table 2, IV B showed a Loading Factor (LF) significantly higher than IV A (entry 2 versus entry 1), highlighting the easier immobilization of IBA in the presence of the longer spacer (that is 1,6-diaminoethane versus 1,2-diaminoethane) [27]. VIII A and VIII B showed LF values of 0.4 and 0.7, respectively, while for VIII C, the iodine LF was found to be 0.3 (Table 1, entries 3–5). The LF values found for IV AB and VIII AC were of the same order of magnitude, and higher than those previously reported for solid reagents based on the immobilization of IBX on both polymer resins and GO [14,20,25]. Moreover, the higher amount of Au with respect to iodine measured for VIII AC proved that the initial linkage of mercapto containing spacers was not quantitative with respect to the Au binding sites available on the support (Table 2, entries 3–5).

3.4. Oxidation of Aromatic Alcohols with IV A–B and VIII A–C

The mechanism of the oxidation of aromatic alcohols with IBX is reported in Scheme 4. The oxygen atom transfer from IBX to the substrate requires the initial addition of the alcohol on activated iodine followed by water elimination and disproportionation with the displacement of the aldehyde [14]. IV AB and VIII AC were applied for the oxidation of a large panel of aromatic alcohols, including benzyl alcohols 16 and phenethyl alcohols 7,8 (Scheme 5, Tables 4 and 5).
The reactions were performed treating the appropriate alcohol (1.0 mmol) with a slight excess of IV AB and VIII AC (1.2 IBX equivalent calculated on the basis of the specific LF value) in EtOAc (10 mL) at 80 °C for 24 h. Tentatively performing the oxidation in other reaction solvents usually applied for IBX transformations (e.g., Dimethyl Sulfoxide (DMSO) and water) were unsuccessful.
Temperatures lower than c.a. 80 °C were not effective, while at temperatures higher than 80 °C, the reagents showed low stability affording only complex mixtures of reaction products. The reactions were analyzed by gas chromatography mass spectrometry (GC–MS) through a comparison with the original standards. Mass–to–charge ratio (m/z) values of aldehydes 916 are reported in Table 3 (the original MS fragmentation spectra are in Figure S1). Under optimal conditions, aromatic aldehydes 916 were detected as the only recovered products aside from unreacted substrates (Table 3 and Table 4). In the case of the oxidation of benzyl alcohol 1, the reaction with commercially available IBX and with IBX supported on polystyrene (sIBX) were performed as references (Table 3, entries 1 and 2).
Homogeneous IBX showed a reactivity higher than the supported reagents in the oxidation of benzyl alcohol 1, probably as a consequence of the diffusional barriers for the access of substrate to active iodine atom, with the only exception of VIII-B, which showed a comparable efficacy (Table 4, entry 1 versus entry 11). On one hand, IV AB and VIII AB oxidized benzyl alcohol 1 to aldehyde 9 in a higher yield with respect to sIBX, suggesting the beneficial role of MWCNTs as support with respect to the organic resin (Table 4 and Table 5). Irrespective of the experimental conditions, VIII-C was totally ineffective in the oxidation of 1, and was not further investigated (Table 5, entry 19). Probably, the low reactivity of VIII-C was ascribable to the detrimental effect of the ester linkage with respect to the amide counterpart on the stability of the Iodine (V) active species [28]. As a general trend, benzyl alcohol derivatives 16 were more reactive than phenethyl alcohols 7,8. Moreover, benzyl alcohol bearing electron donating substituents 25 were more reactive than 1 (Table 4 and Table 5), in accordance with previously reported data focusing on the role of the electron density on the benzylic position in the rate-determining step of IBX-mediated oxidations [14].
The dimension of the spacer also played a significant role, where IV-B and VIII-B bearing the longer spacer chains were the most reactive systems. The effect of the spacer on the reactivity of the supported reagents has been previously investigated, the increase of the length of the chains always being related to the increase of the low energy conformational changes attained by the reagent and to the reduction of the diffusional barrier for the substrates [29]. Finally, Au–MWCNTs based reagents VIII AB were generally more reactive than the MWCNTs counterparts IV AB, most likely due to the increased electron-transfer properties of the support as a consequence of the increased conductance of nanotubes in the Au carbon junctions [30]. The recyclability of supported IBX was evaluated for the more reactive VIII B reagent in the oxidation of benzylic alcohol 1. After the first run, the reagent was recovered by filtration, washed with EtOAc, dried and restored in the active form by treatment with Oxone® and methansulfonic acid. VIII B retained the same reactivity to afford aldehyde 9 in a quantitative yield for at least five successive runs. The absence of leaching of IBX from VIII B was confirmed by testing the oxidative capacity of the organic solution recovered after filtration of its EtOAc solution once maintained at reflux under the same experimental conditions applied for the oxidation. Any oxidation capacity was observed.
Compounds IV B and VIII B retained their morphological structural integrity after the oxidation of alcohol 1, as highlighted by the TEM analysis of the recovered samples (Figure 1, panels B and D, respectively).

4. Conclusions

The preparation of a series of IBX based reagents supported on MWCNTs, as heterogeneous biomimetic systems for the selective oxidation of primary alcohols to the corresponding aldehydes under mild conditions, has been described. Two different types of carbon structures have been investigated, namely oxidized MWCNTs or an alternative, Au-decorated MWCNTs. The immobilization of the iodine active reagent was realized by exploiting the direct formation of an amide linkage with IBX, mediated by different spacer lengths, or through the high binding affinity of sulfur containing linkers in the case of the Au-decorated MWCNTs. In general, the benzyl alcohol derivatives were shown to be more active than the corresponding phenethyl alcohols, thus confirming the prominent role exerted by the electron density on the benzylic carbon in the rate-determining step of the IBX-mediated oxidative process [29]. In accordance with this hypothesis, benzyl alcohol bearing electron donating substituents showed the highest reactivity. The dimension of the spacer incorporated between the IBX fragment and the carbon nanotube surface also played a significant role. Indeed, the reagents bearing the longer spacer chains showed higher LF values and better oxidation performances. Regarding the LF values, the longer spacer may reduce steric hindrances for the IBA-functionalization of MWCNTs, increasing the number of groups involved in the multipoint covalent attachment [31]. Similarly, the better oxidation performance measured in the presence of the longer spacer was in accordance with previously reported data on the role that the spacer length can play for a certain mobility of the active species [32]. Interestingly, Au–MWCNTs based systems behaved as the more reactive reagents, thus justifying their increased electron-transfer properties ascribable to the presence of electroactive Au-carbon joints [33] in comparison with simple MWCNTs counterparts. The novel IBX supported reagents were easily recoverable from the reaction mixture, being successfully used for more runs after a simple reaction with the primary oxidant. These novel reagents can be applied in large scale processes, overcoming drawbacks associated with the use of oxidizing enzymes. Moreover, their metal-free structure, associated with the biocompatibility of MWCNTs, ensures novel reagents high eco-compatibility and low environmental impact.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/8/7/516/s1, Figure S1: original mass spectra of compounds 916; Figure S2: FTIR analysis of oxMWCNTs I; Figure S3: FTIR analysis of II A; Figure S4: FTIR analysis of III A; Figure S5: FTIR analysis of IV A; Figure S6: SEM Back Scattered Electrons (BSE) analysis of VIII B.

Author Contributions

B.M.B., I.A., L.B. prepared the oxidizing solid reagents and synthesized the aldehydes; S.N. and M.F. characterized the catalysts with ICP–MS; M.P. performed and analyzed the XPS measurements; M.C. and R.S. designed the experiments and wrote the paper. A.R.T. performed SEM and TEM analyses.

Acknowledgments

The Filas project MIGLIORA from Regione Lazio and the project PRONAT from CNCCS SCARL are acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Choudhary, V.R.; Dumbre, D.K. Solvent-free selective oxidation of primary alcohols-to-aldehydes and aldehydes-to-carboxylic acids by molecular oxygen over MgO-supported nano-gold catalyst. Catal. Commun. 2011, 13, 82–86. [Google Scholar] [CrossRef]
  2. Christopher, R.G.; Bi-Shun, Z.; SonBinh, T.N. Organic Syntheses by Oxidation with Metal Compounds. J. Am. Chem. Soc. 2006, 128, 12596–12597. [Google Scholar]
  3. Meyer, S.D.; Schreiber, S.L. Acceleration of the Dess-Martin Oxidation by Water Schreiber. J. Org. Chem. 1994, 59, 7549–7552. [Google Scholar] [CrossRef]
  4. Crich, D.; Neelamkavil, S. The fluorous Swern and Corey-Kim reaction: Scope and mechanism. Tetrahedron 2002, 58, 3865–3870. [Google Scholar] [CrossRef]
  5. Sultane, P.R.; Bielawski, C.W. Bielawski. Burgess Reagent Facilitated Alcohol Oxidations in DMSO. J. Org. Chem. 2017, 82, 1046–1052. [Google Scholar]
  6. Noyori, R.; Ohkuma, T. Angew. Asymmetric Catalysis by Architectural and Functional Molecular Engineering: Practical Chemo- and Stereoselective Hydrogenation of Ketones. Chem. Int. Ed. 2001, 40, 40–73. [Google Scholar] [CrossRef]
  7. De Graauw, C.F.; Peters, J.A.; Van Bekkum, H.; Huskens, J. Meerwein-Ponndorf-Verley Reductions and Oppenauer Oxidations: An Integrated Approach. Synthesis 1994, 10, 1007–1017. [Google Scholar] [CrossRef]
  8. Pereira, J.P.C.; Van der Wielen, L.A.M.; Straathof, A.J.J. Perspectives for the microbial production of methyl propionate integratedwith product recovery. Bioresour. Technol. 2018, 256, 187–194. [Google Scholar] [CrossRef] [PubMed]
  9. D’ Acunzo, F.; Baiocco, P.; Fabbrini, M.; Galli, C.; Gentili, P. A Mechanistic Survey of the Oxidation of Alcohols and Ethers with the Enzyme Laccase and Its Mediation by TEMPO. Eur. J. Org. Chem. 2002, 2002, 4195–4201. [Google Scholar] [CrossRef]
  10. Gorrebeeck, C.; Spanoghe, M.; Lanens, D.; Lemière, G.L.; Dommisse, R.A.; Lepoivre, J.A.; Alderweireldt, F.C. Permeability studies on horse liver alcohol dehydrogenase (HLAD) in polyacrylamide gel beads. Recl. Trav. Chim. Pays-Bas 1991, 110, 231–235. [Google Scholar] [CrossRef]
  11. Snijder-Lambers, A.M.; Vulfson, E.N.; Dodema, H.J. Optimization of alcohol dehydrogenase activity and NAD(H) regeneration in organic solvents. Recl. Trav. Chim. Pays-Bas 1991, 110, 226–230. [Google Scholar] [CrossRef]
  12. Liu, R.; Liang, X.; Dong, C.; Hu, X. Transition-Metal-Free:  A Highly Efficient Catalytic Aerobic Alcohol Oxidation Process. J. Am. Chem. Soc. 2004, 126, 4112–4113. [Google Scholar] [CrossRef] [PubMed]
  13. Mulbaier, M.; Giannis, A. The synthesis and Oxidative Properties of Polymer-Supportes IBX. Angew. Chem. Int. Ed. 2001, 40, 23. [Google Scholar] [CrossRef]
  14. Jang, H.-A.; Kim, Y.-A.; Kim, Y.-O.; Lee, S.-M.; Kim, J.W.; Chung, W.; Lee, Y. Polymer-supported IBX amide for mild and efficient oxidation reactions. J. Ind. Eng. Chem. 2014, 20, 29–36. [Google Scholar] [CrossRef]
  15. Verma, N.; Kumar, S.; Ahmed, N. LiBr/β-CD/IBX/H2O-DMSO: A new approach for one-pot biomimetic regioselective ring opening of chalcone epoxides to bromohydrins and conversion to 1,2,3-triketones. Synth. Commun. 2017, 47, 1110–1120. [Google Scholar] [CrossRef]
  16. Kumar, S.; Ahmed, N. β-Cyclodextrin/IBX in water: Highly facile biomimetic one pot deprotection of THP/MOM/Ac/Ts ethers and concomitant oxidative cleavage of chalcone epoxides and oxidative dehydrogenation of alcohols. Green Chem. 2016, 18, 648–656. [Google Scholar] [CrossRef]
  17. Koguchi, S.; Mihoya, A.; Mimura, M. Alcohol oxidation via recyclable hydrophobic ionic liquid-supported IBX. Tetrahedron 2016, 72, 7633–7637. [Google Scholar] [CrossRef]
  18. Zhdankin, V.V. Organoiodine (V) Reagents in Organic Synthesis. J. Org. Chem. 2011, 76, 1185–1197. [Google Scholar] [CrossRef] [PubMed]
  19. Keshavarz, M. Polymer-Supported Hypervalent Iodine as Green Oxidant in Organic Synthesis. Synlett 2011, 16, 2433–2434. [Google Scholar] [CrossRef]
  20. More, J.D.; Finney, N.S. A Simple and Advantageous Protocol for the Oxidation of Alcohols with o-Iodoxybenzoic Acid (IBX). Org. Lett. 2002, 4, 3001–3003. [Google Scholar] [CrossRef] [PubMed]
  21. Kim, Y.H.; Jang, H.S.; Kim, Y.; Ahn, H.; Yeo, Y.; Lee, S.; Lee, Y. Heterogeneous Transition-Metal-Free Alcohol Oxidation by Graphene Oxide Supported Iodoxybenzoic Acid in Water. Synlett 2013, 24, 2282–2286. [Google Scholar]
  22. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Z.; Dubonos, S.V. Electric Field Effect in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [PubMed]
  23. Botta, L.; Bizzarri, B.M.; Crucianelli, M.; Saladino, R. Advances in biotechnological synthetic applications of carbon nanostructured systems. J. Mater. Chem. B 2017, 5, 6490. [Google Scholar] [CrossRef]
  24. Bao, Y.; Pang, H.; Xu, L.; Cui, C.-H.; Jiang, X.; Yan, D.-X.; Li, Z.-M. Influence of surface polarity of carbon nanotubes on electric field induced aligned conductive network formation in a polymer melt. RSC Adv. 2013, 3, 24185. [Google Scholar] [CrossRef]
  25. Parra, E.A.; Blondeau, P.; Crespo, G.A.; Rius, F.X. An effective nanostructured assembly for ion-selective electrodes. An ionophore covalently linked to carbon nanotubes for Pb2+ determination. Chem. Commun. 2011, 47, 2438–2440. [Google Scholar] [CrossRef] [PubMed]
  26. D’Archivio, A.A.; Maggi, M.A.; Odoardi, A.; Santucci, S.; Passacantando, M. Adsorption of triazine herbicides from aqueous solution by functionalized multiwall carbon nanotubes grown on silicon substrate. Nanotechnology 2018, 29, 065701. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Cegłowski, M.; Narkiewicz, U.; Pełech, I.; Schroeder, G. Functionalization of gold-coated carbon nanotubes with self-assembled monolayers of thiolates. J. Mater. Sci. 2012, 47, 3463–3467. [Google Scholar] [CrossRef]
  28. Zhdankin, V.V.; Koposov, A.Y.; Netzel, B.C.; Yashin, N.V.; Rempel, B.P.; Ferguson, M.J.; Tykwinski, R.R. IBX amides: A new family of ipervalent iodine reagents. Angew. Chem. Int. Ed. 2003, 42, 2194–2196. [Google Scholar] [CrossRef] [PubMed]
  29. De Munari, S.; Frigerio, M.; Santagostino, M. Hypervalent iodine oxidants: structure and kinetics of the reactive intermediates in the oxidation of alcohols and 1,2-Diolsby o-iodoxybenzoic acid (IBX) and Dess-Martin periodinane. A comparative 1H-NMR Study. J. Org. Chem. 1996, 61, 9272–9279. [Google Scholar] [CrossRef]
  30. Bayrut, T.; Carlson, J.; Godfrey, B.; Retzlaff, M.S.; Sligar, S.G. Structure, behavior, and manipulation of nanoscale biological assemblies. In Nanostructured Materials and Nanotechnology, 1st ed.; Nalwa, H.S., Ed.; Academic Press: San Diego, CA, USA, 2002; p. 808. ISBN 9780125139205. [Google Scholar]
  31. Khoo, K.H.; Chelikowsky, J.R. Electron transport across carbon nanotube junctions decorated with Au nanoparticles: Density functional calculations. Phys. Rew. B 2009, 79, 205422. [Google Scholar] [CrossRef]
  32. Dos Santos, J.C.S.; Barbosa, O.; Ortiz, C.; Berenguer-Murcia, A.; Rodrigues, C.R.; Fernandez-Lafuente, R. Importance of the Support Properties for Immobilization or Purification of Enzymes. ChemCatChem 2015, 7, 2413–2432. [Google Scholar] [CrossRef] [Green Version]
  33. Román-Martínez, M.C.; De Le Cea, S.M. Heterogenization of homogeneous catalysts on carbon material. In New and Future Developments in Catalysis, 1st ed.; Suib, S.L., Ed.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 55–75. ISBN 9780444538765. [Google Scholar]
Scheme 1. Preparation of IBX supported MWCNTs oxidizing solid reagents IV AB.
Scheme 1. Preparation of IBX supported MWCNTs oxidizing solid reagents IV AB.
Nanomaterials 08 00516 sch001
Scheme 2. Preparation of Au decorated IBX supported MWCNTs oxidizing solid reagents VIII A–B.
Scheme 2. Preparation of Au decorated IBX supported MWCNTs oxidizing solid reagents VIII A–B.
Nanomaterials 08 00516 sch002
Figure 1. TEM images of IV B and VIII B. Panels A and C represent the oxMWCNTs and Au–MWCNTs after the loading procedure of IBX. Panels B and D represent IV B and VIII B recovered after the oxidation of alcohol 1. In panels C and D, the black spot corresponds to the Au particle. (A) IBX-MWCNTs IV B; (B) IBX-MWCNTs IV B after the oxidation of alcohol 1; (C) IBX–Au–MWCNTs VIII B; (D) IBX–Au–MWCNTs VIII B after the oxidation of alcohol 1.
Figure 1. TEM images of IV B and VIII B. Panels A and C represent the oxMWCNTs and Au–MWCNTs after the loading procedure of IBX. Panels B and D represent IV B and VIII B recovered after the oxidation of alcohol 1. In panels C and D, the black spot corresponds to the Au particle. (A) IBX-MWCNTs IV B; (B) IBX-MWCNTs IV B after the oxidation of alcohol 1; (C) IBX–Au–MWCNTs VIII B; (D) IBX–Au–MWCNTs VIII B after the oxidation of alcohol 1.
Nanomaterials 08 00516 g001
Scheme 3. Preparation of Au decorated IBX supported MWCNTs oxidizing solid reagent VIII C.
Scheme 3. Preparation of Au decorated IBX supported MWCNTs oxidizing solid reagent VIII C.
Nanomaterials 08 00516 sch003
Figure 2. XPS of C 1s, O 1s, N 1s, S 2p, Au 4f, and I 3d core level spectra of III B, IV B, VII A, and VIII A compounds.
Figure 2. XPS of C 1s, O 1s, N 1s, S 2p, Au 4f, and I 3d core level spectra of III B, IV B, VII A, and VIII A compounds.
Nanomaterials 08 00516 g002
Scheme 4. General mechanism of oxidation of primary alcohols with IBX.
Scheme 4. General mechanism of oxidation of primary alcohols with IBX.
Nanomaterials 08 00516 sch004
Scheme 5. Oxidation of alcohols 1–8 with IV A–B and VIII A–B.
Scheme 5. Oxidation of alcohols 1–8 with IV A–B and VIII A–B.
Nanomaterials 08 00516 sch005
Table 1. Electron binding energies (eV) of the showed element for the XPS analyzed III B, IV B, VII A and VIII A compounds.
Table 1. Electron binding energies (eV) of the showed element for the XPS analyzed III B, IV B, VII A and VIII A compounds.
ElementReagentAssignments
III BIV BVII AVIII A
N 1s400.5400.5400.5400.5N–H (amide)
S 2p3/2--164.4-C–S–C (sulfide)
168.6168.6C–SOx–C
Au 4f7/2--84.084.0Au-Au
I 3d5/2619.2-619.2-I2
621.5621.5621.5621.5I-O
Table 2. ICP–MS analyses of IV A–B and VIII A–C reagents.
Table 2. ICP–MS analyses of IV A–B and VIII A–C reagents.
EntryCompoundICP-MScLF a
Iodine (%)Gold (%)
1IV A0.03-0.3
2IV B0.21-2.1
3VIII A0.040.380.4
4VIII B0.074.70.7
5VIII C0.032.100.3
a Loading Factor (LF) defined as mmol of iodine per gram of support.
Table 3. Mass–to–charge ratio (m/z) value and the abundance of mass spectra peaks of compounds 9–16.
Table 3. Mass–to–charge ratio (m/z) value and the abundance of mass spectra peaks of compounds 9–16.
Products am/z (%)
Benzaldehyde (9)107 (10) [M+1], 106 (80) [M], 105 (72) (M-1)
4 Methoxy Benzaldehyde (10)136 (20) [M], 135 (69) [M-1], 134 (100) [M-2], 133 (95) [M-3], 132 (77) [M-4], 131 (95) [M-5]
3-4 Dimethoxy Benzaldehyde (11)167 (10) [M+1], 166 (52) [M], 165 (80) [M-1], 164 (92) [M-2], 163 (99) [M-3], 162 (87) [M-4], 161 (50) [M-5], 160 (17) [M-6]
3-4-5 Trimethoxy Benzaldehyde (12)197 (2) [M+1], 196 (85)
4 HydroxyBenzaldehyde (13)124 (2) [M+2], 123 (1) [M+1], 122 (100) [M], 121 (63) [M-1], 120 (10) [M-2]
4 Chlorobenzaldehyde (14)142 (10) [M+2], 141 (12) [M+1], 140 (123) [M], 139 (45) [M-1], 138 (90) [M-2], 137 (50) [M-3], 136 (100) [M-4], 135 (1) [M-5]
4-Hydroxyphenylacetaldehyde (15)136 (100)
Phenylacetaldehyde (16)121 (2) [M+1], 120 (35) [M],
a Mass spectroscopy was performed by using a GC–MS. The peak abundances reported in parentheses.
Table 4. Oxidation of compounds 1–8 with IV A–B a.
Table 4. Oxidation of compounds 1–8 with IV A–B a.
EntrySubstrateOxidantProductYield (%) b
1Benzyl alcohol (1)IBX995
2Benzyl alcohol (1)sIBX c925
3Benzyl alcohol (1)IV A929
44-Methoxy benzyl alcohol (2)IV A1035
53,4-Dimethoxy benzyl alcohol (3)IV A1139
63,4,5-Trimethoxy benzy alcohol (4)IV A1241
74-Hydroxy benzyl alcohol (5)IV A1350
84-Chloro benzyl alcohol (6)IV A1418
9Tyrosol (7)IV A155
10Phenethyl alcohol (8)IV A167
11Benzyl alcohol (1)IV B952
124-Methoxy benzyl alcohol (2)IV B1060
133,4-Dimethoxy benzyl alcohol (3)IV B1162
143,4,5-Trimethoxy benzyl alcohol (4)IV B1263
154-Hydroxy benzyl alcohol (5)IV B1380
164-Chloro benzyl alcohol (6)IV B1445
17Tyrosol (7)IV B1510
18Phenethyl alcohol (8)IV B1615
a The reactions were performed treating the appropriate alcohol (0.1 mmol) with a slight excess of IV A–B (1.2 IBX equivalent calculated on the basis of the specific L.F. value) in EtOAc (1.0 mL) at reflux for 24 h. b The substrate was selectively converted only to the corresponding aldehyde. The yield was evaluated using n-dodecane as the internal standard. The conversion of substrate corresponds to the yield of detected products. c IBX supported on polystyrene.
Table 5. Oxidation of compounds 18 with VIII A–C a.
Table 5. Oxidation of compounds 18 with VIII A–C a.
EntrySubstrateOxidantProductYield (%) b
1Benzyl alcohol (1)IBX995
2Benzyl alcohol (1)sIBX c925
3Benzyl alcohol (1)VIII A938
44-Methoxy benzyl alcohol (2)VIII A1044
53,4-Dimethoxy benzyl alcohol (3)VIII A1148
63,4,5-Trimethoxy benzy alcohol (4)VIII A1250
74-Hydroxy benzyl alcohol (5)VIII A1368
84-Chloro benzyl alcohol (6)VIII A1427
9Tyrosol (7)VIII A159
10Phenethyl alcohol (8)VIII A1610
11Benzyl alcohol (1)VIII B998
124-Methoxy benzyl alcohol (2)VIII B10>99
133,4-Dimethoxy benzyl alcohol (3)VIII B11>99
143,4,5-Trimethoxy benzyl alcohol (4)VIII B12>99
154-Hydroxy benzyl alcohol (5)VIII B13>99
164-Chloro benzyl alcohol (6)VIII B1495
17Tyrosol (7)VIII B1520
18Phenethyl alcohol (8)VIII B1624
19Benzyl alcohol (1)VIII C9<3
20Benzyl alcohol (1)VIII B996 d
a The reactions were performed treating the appropriate alcohol (0.1 mmol) with a slight excess of IV A–B (1.2 IBX equivalent calculated on the basis of the specific L.F. value) in EtOAc (1.0 mL) at reflux for 24 h. b The substrate was selectively converted only to the corresponding aldehyde. The yield was evaluated using n-dodecane as the internal standard. The conversion of the substrate corresponds to the yield of the detected products. c IBX supported on polystyrene. d The recyclability of supported IBX was evaluated for the more reactive VIII B reagent in the oxidation of benzylic alcohol 1, after filtration and treatment with Oxone® and methansulfhonic acid. VIII B retained the same reactivity to afford aldehyde 9 in c.a. quantitative yield for at least five successive runs.

Share and Cite

MDPI and ACS Style

Bizzarri, B.M.; Abdalghani, I.; Botta, L.; Taddei, A.R.; Nisi, S.; Ferrante, M.; Passacantando, M.; Crucianelli, M.; Saladino, R. Iodoxybenzoic Acid Supported on Multi Walled Carbon Nanotubes as Biomimetic Environmental Friendly Oxidative Systems for the Oxidation of Alcohols to Aldehydes. Nanomaterials 2018, 8, 516. https://doi.org/10.3390/nano8070516

AMA Style

Bizzarri BM, Abdalghani I, Botta L, Taddei AR, Nisi S, Ferrante M, Passacantando M, Crucianelli M, Saladino R. Iodoxybenzoic Acid Supported on Multi Walled Carbon Nanotubes as Biomimetic Environmental Friendly Oxidative Systems for the Oxidation of Alcohols to Aldehydes. Nanomaterials. 2018; 8(7):516. https://doi.org/10.3390/nano8070516

Chicago/Turabian Style

Bizzarri, Bruno Mattia, Issam Abdalghani, Lorenzo Botta, Anna Rita Taddei, Stefano Nisi, Marco Ferrante, Maurizio Passacantando, Marcello Crucianelli, and Raffaele Saladino. 2018. "Iodoxybenzoic Acid Supported on Multi Walled Carbon Nanotubes as Biomimetic Environmental Friendly Oxidative Systems for the Oxidation of Alcohols to Aldehydes" Nanomaterials 8, no. 7: 516. https://doi.org/10.3390/nano8070516

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

Article Metrics

Back to TopTop