Next Article in Journal
Application of Three Ecological Assessment Tools in Examining Chromatographic Methods for the Green Analysis of a Mixture of Dopamine, Serotonin, Glutamate and GABA: A Comparative Study
Next Article in Special Issue
Efficient Catalytic Reduction of 4-Nitrophenol Using Copper(II) Complexes with N,O-Chelating Schiff Base Ligands
Previous Article in Journal
Antinociceptive Synergism of Pomegranate Peel Extract and Acetylsalicylic Acid in an Animal Pain Model
Previous Article in Special Issue
Iridaaromatics via Methoxy(alkenyl)carbeneiridium Complexes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Replacement of Volatile Acetic Acid by Solid SiO2@COOH Silica (Nano)Beads for (Ep)Oxidation Using Mn and Fe Complexes Containing BPMEN Ligand

1
CNRS, LCC (Laboratoire de Chimie de Coordination), Université de Toulouse, UPS, INPT, 205, Route de Narbonne, F-31077 Toulouse, France
2
Département de Chimie, Institut Universitaire de Technologie Paul Sabatier, Université de Toulouse, Av. Georges Pompidou, BP 20258, CEDEX, F-81104 Castres, France
3
INPT, École Nationale Supérieure des Ingénieurs en Arts Chimiques et Technologiques, CS 44362, CEDEX 4, F-31030 Toulouse, France
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(18), 5435; https://doi.org/10.3390/molecules26185435
Submission received: 16 July 2021 / Revised: 31 August 2021 / Accepted: 1 September 2021 / Published: 7 September 2021
(This article belongs to the Special Issue Exclusive Feature Papers in Organometallic Chemistry)

Abstract

:
Mn and Fe BPMEN complexes showed excellent reactivity in catalytic oxidation with an excess of co-reagent (CH3COOH). In the straight line of a cleaner catalytic system, volatile acetic acid was replaced by SiO2 (nano)particles with two different sizes to which pending carboxylic functions were added (SiO2@COOH). The SiO2@COOH beads were obtained by the functionalization of SiO2 with pending nitrile functions (SiO2@CN) followed by CN hydrolysis. All complexes and silica beads were characterized by NMR, infrared, DLS, TEM, X-ray diffraction. The replacement of CH3COOH by SiO2@COOH (100 times less on molar ratio) has been evaluated for (ep)oxidation on several substrates (cyclooctene, cyclohexene, cyclohexanol) and discussed in terms of activity and green metrics.

1. Introduction

The synthesis of epoxides/ketones is an interesting research field from the fundamental to the applicative point of view in organic synthesis or catalysis. Indeed, those organic compounds can be obtained using very simple organic oxidants (but quite tedious in the post-treatment procedure) like meta-chloroperbenzoic acid (m-CPBA) [1,2], NaIO4 [3], RCO3H [4,5,6]. They can also be obtained using metal catalysts and the use of an organic solvent is very often required [7,8,9]. It can be the case with several Mo complexes [10,11,12,13,14]. The use of chlorinated solvents such as dichloroethane (DCE), a highly toxic solvent, has to be avoided [15]. In the research group, the processes have been found to be active without organic solvent using complexes with tridentate ligands [16,17,18,19,20] or polyoxometalates (POMs) [21,22,23], giving a first step towards a cleaner process. The oxidant used in this case is tert-butyl hydroperoxide (TBHP) in aqueous solution. In terms of atom economy, the epoxidation reaction could be improved using H2O2 as the oxidant. Selective epoxidation reactions were achieved using (BPMEN)Mn(OTf)2 [24,25,26], (BPMEN)Fe(OTf)2 or (Me2PyTACN)Fe(OTf)2 [27,28,29,30,31,32,33,34,35] as catalysts (BPMEN = N,N′-dimethyl-N,N′-bis(pyridin-2-ylmethyl)ethane-1,2-diamine, Me2PyTACN = 1,4-dimethyl-7-(2-pyridylmethyl)-1,4,7-triazacyclononane), using H2O2 as oxidant in acetonitrile as the organic solvent with high selectivity towards epoxides when acetic acid is added as co-reagent [36,37]. Indeed, by blocking one of the two labile sites on the metal center, the access to cis-diols is not possible [36,37]. Moreover, acting as a proton relay, the carboxylic acid protonates the distal oxygen of the metal-hydroperoxo intermediate, favoring the heterolytic O-O bond cleavage and leading to the clean formation of a metal-oxo compound, an intermediate responsible for the selective oxidation of the olefin into epoxide [37,38]. When BPMEN is used as ligand, a high quantity of acetic acid is used (14 equiv. vs. substrate), with a volume comparable to the one of the organic solvent engaged in the reaction. An elegant way to replace the organic volatile carboxylic acid by recoverable objects could be the use of a solid reagent with COOH pending functions [39,40,41,42]. For this, it was interesting to use the possibility of the functionalization of silica—using trialkoxysilane precursors—to obtain pending acidic functions on silica [43,44,45,46]. Silica was employed previously for different uses, especially to graft, in a covalent way, polydentate ligands and related complexes for catalyzed reactions, or to trap heavy metals for depollution concerns. Those strategies used mainly mesoporous compounds [47,48,49,50,51] but rarely nonporous silica beads. Few examples are related to the replacement of carboxylic function in oxidation reactions catalyzed by Fe or Mn complexes surrounded by tetradentate ligands. Notestein and coworkers reported mono- or di-nuclear Mn complexes of Me3tacn (1,4,7-Trimethyl-1,4,7-triazacyclononane) partially grafted on functionalized mesoporous silica with pendant carboxylic functions. The functions could recover catalyst and replace volatile reagents. Those systems showed interesting results in the oxidation reaction on several substrates [52,53].
In order to find a nonvolatile acidic agent, we used COOH functionalized silica beads instead of acetic acid. To prove the efficiency, the (ep)oxidation reactions were performed with several metal complexes based on BPMEN ligands. Although those metal complexes are not the most efficient for oxygen atom transfer (OAT) reactions, they are advantageous for a proof of concept. Well described in the literature [29,54,55] and with straightforward synthesis, [29] they have well-reported OAT reactivity [55]. The effect of the metal and/or counterion of the catalysts was studied herein. The quantity of COOH functions was evaluated according to the size of the synthesized silica beads. From the results, the green metrics have been used to compare the different methods.

2. Results and Discussion

2.1. Metal Complexes

2.1.1. Synthesis

In order to study the influence of the counter anion during the catalysis and more particularly with the use of the silica beads, three MnII metal complexes with different anions were synthetized according to Figure 1. (L)MnCl2 was obtained in 65% yield by reaction between BPMEN (L) and MnCl2·4H2O in acetonitrile [56]. Similarly, (L)Mn(OTf)2 was obtained in 68% yield [29]. (L)Mn(p-Ts)2 was obtained from (L)MnCl2 via anion metathesis using silver para-toluenesulfonate. Precipitation of AgCl during the reaction confirmed the anion exchange and (L)Mn(p-Ts)2 was isolated in 72% yield.
One FeIII metal complex, [(L)FeCl2](FeCl4), determined by X-ray analysis (vide infra), was obtained in 73% yield by reaction between L and 2 equivalents of FeCl3·6H2O in acetonitrile. It has to be noted that the same reactivity has been observed with other ligands in the literature [57,58].

2.1.2. X-ray Characterization of the Complexes

Suitable crystals for X-ray analysis were obtained for all four metal complexes. The X-ray structures of (L)MnCl2 [56] and (L)Mn(OTf)2 [59] have been previously described in the literature. During the X-ray analysis, the same crystallographic parameters were obtained, confirming the nature of the metal complexes described in Figure 1. Concerning (L)Mn(p-Ts)2 and [(L)FeCl2](FeCl4), their X-ray structures are represented in Figure 2, and principal bond lengths and angles listed in Table 1. Complete data are in Supplementary Materials Tables S1–S3.
In both structures, the metal center is in a distorted octahedral environment. Several ligand-metal-ligand angle values in both metal complexes deviate significantly from the ideal values of a regular octahedron. However, all the angles measured fall in the range found for similar metal complexes in the literature, notably (L)MnCl2 [56] and (L)Mn(OTf)2 [59]. The metal centers are coordinated by the four nitrogen atoms of the L ligand and two anions. In both cases, the two anions are in cis positions and the two pyridine groups of L trans to one another. Consequently, the L ligand folds around the metal center using the cis-α conformation usually observed within this family of aminopyridine ligands.

2.2. Silica Beads

2.2.1. Synthesis

The syntheses of SiO2@COOH (nano)particles were obtained ab initio starting from the synthesis of SiO2 beads—according to a modified Stöber synthesis—using Si(OEt)4 (TEOS) as precursor in presence of aqueous ammonia solution and H2O in alcohol (ethanol or methanol) as solvent (Figure 3) [60]. The influence of solvent, [61] quantity of water, [62,63] concentration of ammonia solution [64] and temperature [65] on the size of silica nanoparticles have already been described in different articles [66]. The size of the particles decreases when solvent polarity increases [67]. Two batches of silica particles were synthesized according to the nature of solvent used during the synthesis. Their reactivity will be compared in several catalyzed oxidation reactions.
The syntheses of SiO2@COOH were performed in two steps (Figure 4). The first step is the functionalization of the surface of the SiO2 nanoparticles by 3-(triethoxysilyl)propionitrile (TESPN) in order to obtain the available nitrile functions SiO2@CN. The terminal nitrile functions were hydrolyzed in a second step into carboxylic ones using H2SO4 (65 wt.%) to obtain the SiO2@COOH beads. All (nano)particles (SiO2, SiO2@CN, SiO2@COOH) were characterized by TEM, DLS, solid NMR and the number of functions grafted quantified by solution 1H NMR.

2.2.2. Characterization

The purpose of two different solvents for the synthesis of the starting SiO2 was to access different beads sizes. Indeed, different sized nonporous silica beads might lead to different specific surfaces (linked to the average diameter of the beads) and might influence the number of grafted functions per gram of silica beads. Thus, objects of different sizes can be added into the reaction media and might change the reactivity and/or the reaction mass efficiency (RME) in the catalyzed oxidation reactions studied herein.
The morphological study of the (nano)particles was done by TEM and DLS to determine their sizes and behaviors in suspension. The proof of the grafting was done using different spectroscopic methods (IR, solid NMR) and the quantification of the grafting through 1H liquid NMR.

Morphological Study

  • Transmission electron microscope (TEM) analysis
From the TEM pictures in the case of the SiO2, SiO2@CN and SiO2@COOH beads (Figure 5), it has been possible to prove the size of the silica beads according to the solvent used. For each step, monodisperse spherical beads have been obtained of around 430−440 nm when produced in ethanol and 62−66 nm when produced in methanol.
  • Dynamic light scattering (DLS) measurements
Monodispersity is an important parameter for SiO2@CN and SiO2@COOH beads, ensuring reproducible catalytic reactions. DLS is another practical and simple method which could determinate the hydrodynamic radius distribution of silica particles.
DLS measurements for SiO2(E), SiO2@CN(E) and SiO2@COOH(E) (E: ethanol) show regular hydrodynamic radii of the particles around 400–450 nm, close to the ones found by TEM, especially because the grafted function thickness is small compared to the bead sizes (Figure 6). The narrow distribution confirmed the relatively monodisperse beads.
In the case of SiO2(M) (M: methanol) beads, for which the size was smaller, the DLS measurements (100 nm for SiO2, 190 nm for SiO2@CN and 68 nm for SiO2@COOH) did not give data in accordance with the observations from TEM. This could be due to some aggregation phenomena or, in the case of SiO2@CN, multilayers of silanes.

Spectroscopic Characterization of the Grafting

  • Infrared spectroscopy
The IR spectra of all silica beads (Figure 7) showed typical vibration bands in accordance with the SiO2 core at 793 cm−1 for Si-O-Si symmetrical vibration, 945 cm−1 for Si-OH, 1060 cm−1 for Si-O-Si asymmetrical ones, 3700 cm−1-2930 cm−1 for -OH in stretching mode. In the case of SiO2@CN vibrations at 2250 cm−1 for CN [68] and 2832 cm−1 for CH stretching mode [69]. The presence of carboxylic functions could be detected, i.e., C=O for SiO2@COOH at 1712 cm−1 [70,71].
The size of the starting SiO2 does give different intensities for the grafted fragments. Indeed, while it is very easy to observe the vibrations assigned to grafted organic part with the SiO2@f(M) beads, it is less obvious in the case of SiO2@f(E). This has to be linked to the grafted functions per size of beads ratio. The smaller the bead is, the “more intense” will be the vibrational pattern of the organic part.
Due to low loading of the grafted functions in the case of SiO2@CN(E) and even lower in SiO2@COOH(E) because of the acid hydrolysis, the vibrations corresponding to functional groups were observed with difficulty from the raw spectra. Those vibrations that could be seen were giving difference spectra between SiO2@CN and SiO2 OR between SiO2@COOH and SiO2, proving the existence of the -CN (Figure 8) and -COOH (Figure 9) functional groups.
  • Solid state NMR
To increase the knowledge about grafting, the multinuclear solid state (CP)MAS NMR (1H, 13C and 29Si) can be investigated. All data have been summarized in Supplementary Materials Table S4. All relevant information will be discussed through nuclei.
The 1H MAS NMR’s very large (and sometimes overlapped) signals are indicative and correspond to different groups on the silica beads, i.e., silanols and physiosorbed water molecules (3.5–5 ppm), EtO (3.3–3.6 ppm), MeO (1.1–1.3 ppm) groups as well as CH2 from the grafted units (0.7–0.9 (Si-CH2), 6.5–6.8 (CH2-N) 4.0–4.1 (CH2)) [72].
The 13C CP-MAS NMR spectra show signals corresponding to the organic functions grafted on SiO2. EtO functions are present in both SiO2 starting beads and after grafting. The signals corresponding to the silane with CN are visible with SiO2@CN, as well as with COOH after the hydrolysis for SiO2@COOH (see Supplementary Materials Table S4 and Figure S1) [72], confirming the grafting and the transformation of the pending function.
The 29Si CP-MAS NMR spectra gave other information (Table S4 and Figure 10). In all spectra, the signals at −93, −101 and −111 ppm corresponding to Q2, Q3 and Q4 respectively (Qn= Si(OSi)n(OH)4−n) are in accordance with SiO2 core [73,74]. The grafting was proved by two signals at around −60 and −70 ppm (T2 and T3) [75]. A change in the proportion of the signals was observed from SiO2 to SiO2@CN and from SiO2@CN to SiO2@COOH, the trend being identical with the starting SiO2(M) and SiO2(E) beads. Since CP MAS could not be used to quantify the Qn, the deconvolutions were performed on MAS spectra (Figure S2). The intensity distribution is summarized in Table S4.
The solid-state NMR showed that the SiO2 beads contain some ethoxy functions (although dried under vacuum) and those functions remain even when the grafting occurs. 29Si NMR spectra exhibit a qualitative change of the silicon core with the grafted functions. In order to use those beads in a precise and quantitative manner, it was important to quantify the grafted functions at the surface through different parameters.
  • Quantification by 1H NMR in solution
When an analyzed sample is simple or pure, elemental analysis (EA) can give accurate information. In the case of the presented silica beads, the system—as shown by multinuclear MAS NMR—is more complex and EA would not give reliable results. One elegant method has been developed [40], considering that, in a very alkaline medium, silica can be transformed into silicates maintaining the integrity of the organic fragments that can be easily quantified by 1H solution NMR, using an internal standard (benzoic acid herein, stable and soluble in very basic solution as benzoate).
Thus, a mass of sample silica beads was dissolved in strong alkaline deuterated aqueous solution (pH ≈ 13) and analyzed by 1H NMR using a mass of internal standard, giving a number of moles of functions per gram of silica beads (all beads, i.e., SiO2, SiO2@CN and SiO2@COOH).
The signals corresponding to ethanol and methanol are related to the alkoxy functions present on beads, from TEOS to TESPN (Figure 11). All the other CH2 signals are related to the non-alkoxy part of TESPN and the corresponding oxidized one. The 1H NMR shifts have been presented in Table S5.
The number of functions n(f) has been calculated based on 1H NMR integrations I(f) relatively to I(ref) from a known mass of internal standard, m(ref) (Table 2). With n(f), the density of f functions per mass of sample ρ(f) was defined according to the mass of SiO2 sample (mS) using Equation (1).
ρ f = n f m S = I f m S · m ref M ref · 1 I ref
The results showed that -OEt fragments were present on starting SiO2, with a higher content per gram of sample with SiO2(M) beads (smaller size) [76,77]. The functionalization was, for the same reason, higher per gram of sample in the case of SiO2@CN(M). From SiO2@CN to SiO2@COOH, the hydrolysis removed a substantial part of the “grafted” functions, certainly destroyed/removed by concentrated sulfuric acid.
  • Determination of function coverage of functionalized silica beads
Using several techniques, it is possible to calculate the function coverage on silica cores, an important parameter within the catalytic part. The parameter μ(f), defined in the number of groups per nm2, could be determined by Equation (3) [23,40]. The ρ’(f) parameter does correspond to the functions grafted on a silica core (Figure 12 and Equation (2)) and is calculated from ρ(f). The average radius of the SiO2 beads (rcore) is deduced from the TEM measurements. μ(f) was calculated with a core mass (mcore) of 1 g.
ρ f = n f m core = ρ f 1 ρ f . M S i l a n e   .
The parameter μ(f) is the number of molecules n(f) grafted on 1 g of the sample surface ΣScore (in nm2). From the SiO2 radii found in TEM measurements, Equation (3) can be written as follows:
μ f = ρ f . r core . ρ SiO 2 3.10 + 21 × N A  
Using Equation (3), coverage by CN and COOH fragments have been calculated (Table 3). Concerning the SiO2@CN, the μ(CN) value is very high (>17) and seems to confirm a multilayer deposition. The μ(COOH) values around 3 for SiO2@COOH are in agreement with what is expected with monolayers.

2.3. Catalysis

The BPMEN-related complexes were tested on three different substrates and two different co-reagents, CH3COOH (in order to use the results as reference) or SiO2@COOH. The catalytic study presented herein will be divided according to the substrates.
The complexes were tested as homogenous catalysts under the classical conditions (using acetic acid as co-reagent) and the influence of the metal and anion was studied. The reactivity was compared with the processes using SiO2@COOH beads or acetic acid. These complexes were tested in olefin epoxidation and alcohol oxidation. For this reason, cyclooctene (CO) was chosen as model substrate for epoxidation, while the (ep)oxidation of cyclohexene (CH) and oxidation of cyclohexanol (CYol) were studied for their potential applied interest towards the synthesis of adipic acid, both being starting reagents in different processes [31,32,33,34,35,78,79].
Reaction under homogeneous conditions was previously described [31,80]. To prevent H2O2 disproportionation [81] and Fenton reaction [82], H2O2 was slowly added at 0°C for two hours [83] (especially in the case of Fe complex) [84] using CH3CN as solvent. The cat/substrate/H2O2/CH3COOH ratio of 1/100/150/1400 was followed. The reactions were stopped after 3 h and analysed by GC-FID using acetophenone as an internal standard.

2.3.1. Oxidation of Cyclooctene

Cyclooctene (CO) was used as the model since the substrate is known to give the corresponding cyclooctene oxide (COE) with high selectivity. To prove the need of carboxylic function as co-reagent in this catalysis, some tests with complexes were done in the absence and presence of co-reagent (Table 4). While no CO conversion was observed with [(L)FeCl2](FeCl4), all (L)MnX2 complexes (X = Cl, OTf, p-Ts) were poorly active, showing the necessity of a carboxylic co-reagent. All complexes were tested in the presence of a co-reagent, acetic acid or SiO2@COOH (taking into account the bead sizes) under identical experimental conditions.
In the presence of a co-reagent (Figure 13), all catalysts could achieve CO conversion, the best conditions being in the presence of acetic acid for manganese complexes, while the conversion was better in the presence of SiO2@COOH with the iron complex (Table 4 and Figure 14). The lower conversion in the presence of SiO2@COOH beads for manganese complexes seems to be due to the heterogeneous character of the reaction. COE was the only product observed by GC-FID. The low selectivity towards COE in the presence of (L)MnX2 (X = OTf, p-Ts) and [(L)FeCl2](FeCl4) might be due to the formation of cyclooctanediol and the subsequent opening ring reaction conducting to suberic acid [85,86]. Those two products could not be observed by GC-FID using the method developed herein.
Using CH3COOH as the co-reagent with a cat/CH3COOH ratio of 1:1400 (Table 4 and Figure 14), the results for the complexes (L)MnX2 (X = Cl, OTf) were similar to those described [29]. The manganese complexes (L)MnX2 (X = Cl, OTf, p-Ts) gave almost complete CO conversion. However, the selectivity towards COE with X = OTf and p-Ts around 60% was lower than X = Cl (81%). It can be concluded that the anion has an influence on the selectivity towards COE. It might be due to the basicity of the anion, the chloride being the more inert. As pointed out previously, the ring opening might occur in presence of acid/base, and it was certainly what happened here. However, diminishing the cat/CH3COOH ratio to 1:14 for (L)MnCl2 gave similar results to the ones observed in the absence of acetic acid, underlying the necessity of a huge excess of co-reagent to achieve high conversion and selectivity with complexes based on BPMEN ligand.
Very interestingly, using SiO2@COOH beads as co reagents with a cat/COOH ratio of 1:14, the conversion of CO was observed, proving the positive effect of the silica beads functionalized with COOH even with a relatively low amount of COOH functions in the reactional mixture In addition, the use of SiO2@COOH beads as co-reagents gave in the case of the manganese complexes a reverse effect (Table 4 and Figure 13) than the one observed with acetic acid. Indeed, the conversion follows the X order p-Ts > OTf > Cl, with a selectivity towards COE in favor of the triflate, followed by the p-Ts and finally the chloride salt. The effect of the bead size is negligible in the case of the two more active complexes ((L)MnX2 (X = OTf, p-Ts)) while a stronger difference is observed with the chloride salt, giving lower selectivity towards COE.
Concerning the iron complex, a moderate conversion and a low selectivity were observed in the presence of CH3COOH. With silica beads, higher conversions were obtained and the selectivities were similar to the ones with CH3COOH.

2.3.2. Oxidation of Cyclohexene

The cyclohexene (CH) is a very interesting substrate as a starting material for the synthesis of adipic acid [22,79]. In comparison to CO, the (ep)oxidation of CH is more complex. Indeed, according to the nature of the metal used within the reaction, two oxidations are possible: allylic oxidation on sp3 C-H bonds and epoxidation on C=C double bond [87]. Other possible water additions and/or subsequent oxidation give a complex mixture.
Cyclooctene oxide (CHO), cyclohexanediol (CHD), cyclohexene-1-ol (CHol) and cyclohexen-1-one (CHone) are the most common observed products (see Figure 15). The conversion of CH, the selectivity towards the products and TON have been compiled (Table 5, and Figure 16).
All the manganese complexes (L)MnCl2 (X = Cl, OTf, p-Ts) exhibited high CH conversion in the presence of CH3COOH and the analysed products are anion-dependent. While X = Cl gave exclusively CHO with a relatively good selectivity (89%), the complexes with X = OTf and p-Ts gave a small quantity of CHD and CHone. When SiO2@COOH beads were used instead of acetic acid, the CH conversions were lower, CHO being the only product detected with X = OTf and p-Ts. (L)MnCl2 showed a part of ring opening (presence of CHD) with SiO2@COOH(E) beads and allylic oxidation (presence of CHol and CHone) with the SiO2@COOH(M). From those observations, it seems that the presence of CH3COOH or SiO2@COOH have reverse effects in terms of selectivity according to the nature of the anion of the Mn complex. This has certainly to be linked to the mechanism occurring between the manganese complex and the co-reagent linked to the nature of the interaction between the anion and the “MnL” part.
With the [(L)FeCl2]FeCl4 complex, the mechanism seems to be radically different since the reaction with CH3COOH as co-reagent gave hardly any product (although a slight conversion was observed). Surprisingly, the use of SiO2@COOH did improve the CH conversion but not in a selective way since the products originating from epoxidation and allylic oxidation were observed in almost equal quantities.

2.3.3. Oxidation of Cyclohexanol

The cyclohexanol (CYol) is also a very interesting substrate as a starting material of the KA oil (KA oil = ketone-alcohol oil) used for the synthesis of adipic acid [88,89]. In addition, compared to the oxidation of CH, oxidation of CYol gives only one product, i.e., cyclohexanone (CYone) (see Figure 17). Catalyzed cyclohexanol oxidation followed the same procedure as CO and CH and results have been compiled in Figure 18 and Table 6.
With all complexes, in the presence of CH3COOH, the conversion of CYol was high and selective towards CYone [90,91]. (L)Mn(OTf)2 and (L)Mn(p-Ts)2 complexes were more active than (L)MnCl2. Due to the lability of OTf and p-Ts anions, the coordination site in (L)Mn(OTf)2 and (L)Mn(p-Ts)2 was more accessible than for (L)MnCl2. As a consequence, the access to the metal center for peroxide and carboxylic function might be favored. Due to the heterogeneous nature of the SiO2@COOH reagent, the conversion was lower in all cases. Some differences appeared in terms of selectivity, due to the nature of the anion within the complexes (in the case of the manganese complexes) and/or to the nature of the metal in the case of the iron complex. Notably, selectivity was drastically diminished for the iron complex in the presence of SiO2@COOH.

2.4. Green Metrics

The use of SiO2@COOH is interesting in terms of the material recovery parameter. Indeed, the studied parameter between all tests has been the replacement of acetic acid by the silica beads, and it has to be pointed out that the number of carboxylic functions is lower with the beads (from a factor 100). Some green metrics could be considered within this process [92]. The recovery of by-products (water, acetic acid and excess H2O2) would require more energy than the distillation of acetonitrile and the filtration/centrifugation of the silica beads. The difference lies, thus, in the non recovered waste.
Considering the several green metrics, the atom efficiency (AE) and stoichiometric factors (SF)—being identical for all the studied reaction—were not added in the comparisons. The yield, the MRP and RME have been graphically presented. It can be seen that most reactions have lower yields when SiO2@COOH is used but with a slightly better RME (the mass of beads is lower than the mass of acetonitrile, even the bigger ones). The MRP is in each case in favor to SiO2@COOH. (Figure 19, Figure 20 and Figure 21). Those results represent a proof of a concept of a cleaner process.

3. Materials and Methods

3.1. Materials

All manipulations were carried out under air. Distilled water was used directly from a Milli-Q purification system (Millipore, Burlington, MA, USA). Acetonitrile, ethanol, methanol (synthesis grade, Aldrich) were used as solvents as received. Tetraethyl orthosilicate (TEOS, 98% Aldrich, St. Louis, MI, USA), ammonium hydroxide solution (25%, Aldrich), 3-(Triethoxysilyl)propionitrile (97%, Aldrich), cis-cyclooctene (95%, Alfa Aesar, Karlsruhe, Germany), cyclooctene oxide (99%, Aldrich), cyclohexene (99%, Acros), cyclohexene oxide (98%, Aldrich), 2-cyclohexen-1-ol (95%, TCI, Tokyo, Japan)), 2-cyclohexen-1-one (96%,TCI), cis-1,2-cyclohexanediol (99%, Acros, Geel, Belgium), cyclohexanol (99%, Alfa Aesar), cyclohexanone (99.8%, Acros) and TBHP (70% in water, Aldrich) were used as received.

3.2. Methods

3.2.1. X-ray Structural Analyses

A single crystal of each compound ((L)Mn(p-Ts)2 and [(L)FeCl2](FeCl4)) was mounted under inert perfluoropolyether on the tip of a glass fiber and cooled in the cryostream of a Bruker Nonius CAD4 APEXII diffractometer. The structures were solved by using the integrate space-group and crystal structure determination SHELXT software [93] and refined by least squares procedures on F2 using SHELXL-2014 [94]. The crystal and refinement parameters of all compounds are collected in Table S1 and the full list of bond distances and angles provided in Supplementary Materials Tables S2 and S3. All H atoms attached to carbon were introduced in calculation in idealized positions and treated as riding models. The drawing of the molecules was realized with the help of ORTEP32 [95,96]. CCDC 1959449 (for (L)Mn(p-Ts)2) and 1959450 (for [(L)FeCl2](FeCl4)) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre.

3.2.2. Dynamic Light Scattering

Preparation sample: in order to be able to obtain repetitive and correct data analysis, particle samples were prepared at 0.1 wt.% in water. A sonication of the particles suspension was made before DLS analysis for 5 min at 350 W (FB705 Fisherbrand Ultrasonic Processor), facilitating the dispersion of silica particles. Hydrodynamic diameters of the particles in suspension were obtained with a ZetaSizer Nano-ZS (Malvern Instruments Ltd.). This equipment uses a laser (He-Ne at λ = 633 nm, under voltage of 3 mV) and the detector is located at 173° to analyse the scattered intensity fluctuations. A portion of 10 mg of particles was dispersed in 20 mL of water with the ultrasonic processor 40 (5 min, 350 W) prior to the measurement performed at a temperature of 25 °C.

3.2.3. TEM

Particle morphology was performed with a JEOL JEM1011 transmission electron microscope equipped with 100 kV voltage acceleration and tungsten filament (Service Commun de Microscopie Electronique TEMSCAN, Centre de Microcaractérisation Raimond Castaing, Toulouse, France). A drop of sonicated particle solution (0.1 wt.% in ethanol) was disposed on a formvar/carbon-coated copper grid (400 mesh) and dried in air for 48 h.

3.2.4. Infrared Spectroscopy

Fourier Transform infrared (FTIR) spectra were recorded by Spectrum two—PerkinElmer.

3.2.5. Solid State NMR

NMR experiments were recorded on Bruker Avance 400 III HD spectrometers operating at magnetic fields of 9.4 T. Samples were packed into 4 mm zirconia rotors. The rotors were spun at 8 kHz at 293 K. 1H MAS was performed with DEPTH pulse sequence and a relaxation delay of 3 s. For 29Si MAS single pulse experiments, small flip angle of 30° was used with recycle delays of 60 s. 13C CP and 29Si CP MAS spectra were recorded with a recycle delay of 2 s and contact times of 3 ms and 4 ms, respectively. Chemical shifts were referenced to TMS. All spectra were fitted using the DMfit software.

3.2.6. Solution NMR

1H-NMR and 13C-NMR spectra were recorded on Bruker NMR III HD 400 MHz spectrometers, 400 MHz for 1H-NMR, and 101 MHz for 13C-NMR.

3.2.7. Elemental Analysis

Elemental analyses were performed by the microanalysis service of the LCC.

3.2.8. Centrifugation

The silica beads were collected by centrifugation on a Fisher 2-16P with 11192 rotor (Max. rpm 4500, Sigma).

3.2.9. Gas Chromatography

The catalytic reactions were followed by gas chromatography on an Agilent 7820A chromatograph equipped with an FID detector, a DB-WAX capillary column (30 m × 0.32 mm × 0.5 μm) and autosampler. Authentic samples of reactants (cyclooctene, cyclohexene, cyclohexanol) and some potential products (cyclooctene oxide, cyclohexene oxide, 2-cyclohexen-1-ol, cis-1.2-cyclohexanediol, 2-cyclohexen-1-ol, and cyclohexanone) were used for calibration. The conversion and the formation were calculated from the calibration curves (r2 = 0.999) and an internal standard.

3.2.10. Quantification of the Number of Functions per Gram of Grafted Silica through 1H NMR in Solution

A sample of 7 mg of SiO2@R (R= CN, COOH) was added to 4 mL of D2O/NaOH solution (pH ≈ 13) in an NMR tube. The mixture was heated until the powder completely dissolved. A known amount of benzoic acid (ca. 4 mg) was added as internal standard. Then the NMR proton data were collected immediately.

3.3. Synthesis of Metal Complexes

3.3.1. (L)MnCl2

According to ref [56] MnCl2,4H2O (0.48 g, 2.4 mmol) was added to a solution of L (0.54 g, 2 mmol) in 3 mL of acetonitrile. The mixture was stirred at room temperature for 15 h and the solvent was removed under vacuum. The grey powder obtained was washed twice with diethyl ether and after recrystallization by diffusion of diethyl ether into a solution of the product in an acetonitrile-ethanol mixture, (L)MnCl2 (0.52 g, 65% yield) was obtained as a white powder.
Anal. Calc. for C16H22Cl2MnN4·0.5EtOH: C, 48.70; H, 6.01; N, 13.36. Found: C, 49.02; H, 5.98; N, 13.40.

3.3.2. (L)Mn(OTf)2

According to ref [29], Mn(OTf)2 (0.875 g, 2.4 mmol) was added to a solution of L (0.54 g, 2 mmol) in 3 mL of acetonitrile. The mixture was stirred at room temperature for 15 h and the solvent was removed under vacuum. The light grey powder obtained was washed twice with diethyl ether and after recrystallization by diffusion of diethyl ether into a solution of the product in acetonitrile, (L)Mn(OTf)2 (0.85 g, 68% yield) was obtained as a white powder.
Anal. Calc. for C18H22F6MnN4O6S2: C, 34.68; H, 3.56; N, 8.99. Found: C, 34.68; H, 3.42; N, 8.95.

3.3.3. (L)Mn(p-Ts)2

A solution of Ag(p-Ts) (1.34 g, 4.8 mmol) in 5 mL of H2O was added to a solution of (L)MnCl2 (0.79 g, 2 mmol) in 5 mL of H2O and the mixture was stirred at room temperature for 15 h. After removal of the AgCl precipitate by filtration, the solvent was removed under vacuum. Recrystallization of the crude product in absolute ethanol afforded (L)Mn(p-Ts)2 (0.96 g, 72% yield) as a grey solid.
Anal. Calc. for C30H36MnN4O6S2: C, 53.97; H, 5.43; N, 8.39. Found: C, 53.82; H, 5.50; N, 8.36.

3.3.4. [(L)FeCl2](FeCl4)

FeCl3,6H2O (1.08 g, 4 mmol) was added to a solution of L (0.54 g, 2 mmol) in 5 mL of acetonitrile. After 15 min, a red precipitate appeared and the mixture was stirred for 15 h at room temperature. After filtration of the red solid recrystallization in CH3CN afforded [(L)FeCl2](FeCl4) (0.93 g, 73% yield) as a red solid.
Anal. Calc. for C16H22Cl6Fe2N4: C, 32.31; H, 3.73; N, 9.42. Found: C, 32.39; H, 3.16; N, 9.33.

3.4. Synthesis of Silica Particles

3.4.1. SiO2 Particles in EtOH (SiO2(E))

According to ref [64], 72 mL (4 mol) of H2O, 60 mL of ammonic solution (28% wt) were mixed in 630 mL (10.79 mol) of absolute ethanol at room temperature. A measure of 40 mL (0.18 mol) of tetraethylorthosilicate (TEOS) was added to the solution. A white suspension appeared. The mixture was stirred at 50 °C for 6 h. Then the solid was washed with absolute ethanol 5 times and collected by centrifugation. SiO2(E) particles were dried under vacuum at 120 °C overnight. A white powder was obtained.
SiO2(E): 1H NMR (400 MHz, D2O/NaOH-Benzoic acid) δ 7.57 (m, 2H, Ar-H), 7.21 (m, 3H, Ar-H), 3.31 (q, J = 7.1 Hz, 0.3H, CH2), 0.86 (t, J = 7.1 Hz,.0.43H, CH3). Anal. Found: C, 1.09; H, 0.67. 29Si CP MAS-NMR: -93.3 ppm (Q2), −101.9 ppm (Q3), −111.8 ppm (Q4). 13C CP MAS-NMR: 58.0 ppm (CH2O), 16.9 ppm (CH3). IR (ATR, ν(cm−1)): 3710-2935 (OH), 1059 (Si-O-Si), 949 (Si-OH), 790 and 438 (Si-O-Si).

3.4.2. SiO2@CN(E) Particles

According to ref [68], a measure of 10 g of SiO2(E) particles was mixed with 25 mL of TESPN (0.11 mol) in 150 mL of toluene under stirring at 110°C for 6 days. The powder was washed 5 times with toluene, collected by centrifugation and dried under vacuum at 120°C overnight to obtain SiO2@CN(E) as a white powder.
1H NMR (400 MHz, D2O/NaOH-Benzoic acid) δ 7.66 (m, 2H, Ar-H), 7.29 (m, 3H, Ar-H), 3.42 (q, J = 7.1 Hz, 0.36H, CH2), 2.15 (m, 0.23H, CH2), 0.96 (t, J = 7.1 Hz, 0.54H, CH3), 0.54 (m, 0.24H, CH2). 29Si CP MAS-NMR: −62.2 ppm (T2), −70.4 ppm (T3), −92.8 ppm (Q2), −101.9 ppm (Q3), −111.9 ppm (Q4). 13C CP MAS-NMR: 120.9 ppm (CN), 60.2 ppm (CH2O), 58.1 ppm (CH2O), 16.4 ppm (CH3), 10.6 ppm (CH2Si), 8.8 ppm (CH2Si). IR (ATR, ν(cm−1)): 3712–2937 (OH), 2248 (CN), 1073 (Si-O-Si), 943 (Si-OH), 795 and 442 (Si-O-Si). ρ(CN) = 0.29 mmol/g. μ(CN) = 20.6 functions/nm2

3.4.3. SiO2@COOH(E) Particles

According to ref [39], a measure of 5 g of SiO2@CN(E) was added to 50 mL of H2SO4 (65% wt, 0.52 mol) and the solution was heated at 150 °C under stirring for 4 h. A grey powder was found in suspension. Then the powder was washed with H2O until pH = 7. The product was collected by centrifugation and was dried under vacuum at 120°C. A light grey powder of SiO2@COOH(E) was obtained.
1H NMR (400 MHz, D2O/NaOH-Benzoic acid) δ 7.58 (m, 2H, Ar-H), 7.21 (m, 3H, Ar-H), 3.33 (q, J = 7.1 Hz, 0.16H, CH2), 1.91 (m, 0.02H, CH2), 0.87 (t, J = 7.1 Hz, 0.23H, CH3), 0.54 (m, 0.03H, CH2). 29Si CP MAS-NMR: -59.6 ppm (T2), −68.7 ppm (T3), −92.8 ppm (Q2), −101.9 ppm (Q3), −111.7 ppm (Q4).13C CP MAS-NMR: 60.0 ppm (CH2O), 58.8 ppm (CH2O), 16.6 ppm (CH3). IR (ATR, ν(cm−1)): 3709–2933 (OH), 1737–1716 (C=O), 1073 (Si-O-Si), 943 (Si-OH), 794 and 446 (Si-O-Si). ρ(COOH) = 0.04 mmol/g. μ(COOH) = 2.8 functions/nm2.

3.4.4. SiO2 Nanoparticles in Methanol (SiO2(M))

A measure of 72 mL (4 mol) of H2O and 60 mL of ammonic solution (28 wt.%) were mixed in 630 mL (15.57 mol) of methanol at room temperature. A measure of 40 mL (0.18 mol) of tetraethyl orthosilicate (TEOS) was added into the solution. A suspension of a white solid appeared. The mixture was stirred at 50 °C for 6 h. The solid was washed with absolute ethanol 5 times, collected by centrifugation and dried under vacuum at 120°C overnight. A white powder of SiO2(M) was obtained.
1H NMR (400 MHz, D2O/NaOH-Benzoic acid) δ 7.62 (m, 2H, Ar-H), 7.25 (m, 3H, Ar-H), 3.06 (s, 0.04H, CH3). 29Si CP MAS-NMR: −93.3 ppm (Q2), −101.9 ppm (Q3), −111.7 ppm (Q4). 13C CP MAS-NMR: 58.2 ppm (CH2O), 16.7 ppm (CH3). IR (ATR, ν(cm−1)): 3732–2850 (OH), 1062 (Si-O-Si), 945 (Si-OH), 784 and 443 (Si-O-Si).

3.4.5. SiO2@CN(M) Nanoparticles

A measure of 10 g of SiO2(M) was mixed with 25 mL of TESPN (0.11 mol) in 150 mL of toluene at 110°C under stirring for 6 days. The solid was washed 5 times with toluene, collected by centrifugation and dried under vacuum at 120 °C overnight. A white powder of SiO2@CN(M) was obtained.
1H NMR (400 MHz, D2O/NaOH-Benzoic acid) δ 7.66 (m, 2H, Ar-H), 7.30 (m, 3H, Ar-H), 3.43 (q, J = 7.1 Hz, 3.16H, CH2), 3.12 (s, 0.06H, CH3), 2.20 (m, 1.98H, CH2), 0.96 (t, J = 7.1 Hz, 4.76H, CH3) 0.54 (m, 2.02H, CH2). 29Si CP MAS-NMR: −64.7 ppm (T2), −70.5 ppm (T3), −93.3 ppm (Q2), −102.4 ppm (Q3), −111.7 ppm (Q4). 13C CP MAS-NMR: 121.0 ppm (CN), 59.9 ppm (CH2O), 16.8 ppm (CH3), 10.5 ppm (CH2Si), 8.8 ppm (CH2Si). IR (ATR, ν(cm−1)): 3721-2903 (OH), 2253 (CN), 1065 (Si-O-Si), 941 (Si-OH), 788 and 425 (Si-O-Si). ρ(CN) = 1.40 mmol/g. μ(CN) = 16.6 functions/nm2

3.4.6. SiO2@COOH(M) Nanoparticles

A measure of 5 g of SiO2@CN(M) was added to 50 mL of H2SO4 (65% wt, 0.52 mol) and the solution was heated at 150 °C under stirring for 4 h. A grey powder was found in suspension. The powder was washed with H2O until pH = 7. The product was collected by centrifugation and was dried under vacuum at 120 °C. A light grey powder of SiO2@COOH(M) was obtained.
1H NMR (400 MHz, D2O/NaOH-Benzoic acid) δ 7.66 (m, 2H, Ar-H), 7.29 (m, 3H, Ar-H), 3.42 (q, J = 7.1 Hz, 0.03H, CH2), 3.12 (s, 0.03H, CH3), 1.99 (m, 0.12H, CH2), 1.02 (t, J = 7.1 Hz, 0.04H, CH3), 0.46 (m, 0.13H, CH2). 29Si CP MAS-NMR: −58.8 ppm (T2), −68.4 ppm (T3), −91.9 ppm (Q2), −101.8 ppm (Q3), −111.6 ppm (Q4). 13C CP MAS-NMR: 177.9 ppm (COOH), 59.9 ppm (CH2O), 49.5 ppm (CH2O), 16.7 ppm (CH3), 6.7 ppm (CH2Si).IR (ATR, ν(cm−1)): 3709–2852 (OH), 1717 (C=O), 1046 (Si-O-Si), 932 (Si-OH), 785 and 450 (Si-O-Si). ρ(COOH) = 0.31 mmol/g. μ(COOH) = 3.2 functions/nm2.

3.5. Catalytic Experiments

3.5.1. General Procedure of Catalysis with CH3COOH

A measure of 1 mmol of substrate (CO, CH. CYol), 0.84 g (14 mmol or 0.14 mmol) of CH3COOH, 0.01 mmol of complexes ((L)MnCl2, (L)Mn(OTf)2, (L)Mn(p-Ts)2, [(L)FeCl2](FeCl4)) and some drops of an internal standard (acetophenone) were mixed in 2 mL of CH3CN at room temperature. A measure of 0.13 mL of H2O2 (35 wt.% in H2O) diluted into 0.87 mL of CH3CN was slowly added into the mixture for 2 h at 0 °C. The mixture was left for 1 h at 0 °C.

3.5.2. General Procedure of Catalysis with SiO2@COOH

A measure of 1 mmol of substrate (CO, CH, CYol), 300 mg of SiO2@COOH(E) (13.5 mg for SiO2@COOH(M) (0.14 mmol of carboxylic function), 0.01 mmol of complexes ((L)MnCl2, (L)Mn(OTf)2, (L)Mn(p-Ts)2, [(L)FeCl2](FeCl4)) and some drops of an internal standard (acetophenone) were mixed in 2 mL of CH3CN at room temperature. A measure of 0.13 mL of H2O2 (35 wt.% in H2O) diluted in 0.87 mL of CH3CN was slowly added to the mixture for 3 h at 50 °C. Then the mixture was left at 60 °C for 2 h.

4. Conclusions

It has been possible to replace acetic acid with silica beads with carboxylic functions in the reaction of the epoxidation of olefins. The study showed lower activity with the silica beads in the case of cyclooctene and cyclohexene oxidation with manganese complexes and selectivity seemed to be linked to the nature of the ion of the complex. With cyclohexene, the activity with the beads was higher relatively to cyclooctene. However, for the Fe complex, the beads were more active than acetic acid. With cyclohexanol, the process worked much better with acetic acid. The size of the bead seemed to have no relevant effect in terms of efficiency, except that the quantity of carboxylic functions brought into the reaction was 100 times less than the quantity of acetic acid. It should be noted that under a lower quantity of acetic acid, the reaction did not work. Although less active, this method is the first step towards the replacement of an organic volatile reagent.

Supplementary Materials

The following are available online, Table S1: Crystal data. Table S2: Bond lengths [Å] and angles [°] for (L)Mn(p-Ts)2. Table S3: Bond lengths [Å] and angles [°] for [(L)FeCl2](FeCl4). Table S4: Relevant solid-state NMR data. Table S5: 1H NMR chemical shifts (in ppm) observed with SiO2, SiO2@CN and SiO2@COOH in D2O/NaOH (pH=13) solution. Figure S1: 13C MAS NMR spectra of SiO2 (bottom), SiO2@CN (middle) and SiO2@COOH (top) for beads from SiO2 beads produced in EtOH (left) and MeOH (right). Figure S2: 29Si MAS NMR spectra of SiO2 (top) SiO2@CN (middle), SiO2@COOH (bottom) from SiO2 beads produced in EtOH (left) and MeOH (right).

Author Contributions

Conceptualization, D.A. and P.G.; methodology, D.A. and P.G.; validation, Y.W., P.G., F.G., J.-C.D. and D.A.; formal analysis, Y.W., J.-C.D., F.G. and D.A.; investigation, Y.W., P.G. and D.A.; resources, Y.W., D.A., P.G. and F.G.; writing—original draft preparation, Y.W., P.G. and D.A.; writing—review and editing, Y.W., F.G., P.G. and D.A.; supervision, P.G. and D.A.; project administration, D.A.; funding acquisition, D.A., P.G. and F.G. All authors have read and agreed to the published version of the manuscript.

Funding

The research leading to these results has received co-funding from Région Languedoc Roussillon Midi-Pyrénées (Région Occitanie), IUT A Paul Sabatier and “Syndicat Mixte de la Communauté d’Agglomération Castres-Mazamet” under grant agreement no. 15066786 for the Y. W. Ph.D. fellowship.

Data Availability Statement

Data is contained within the article or Supplementary Material.

Acknowledgments

The authors acknowledge LCC-CNRS for elemental analyses, solid state NMR and TEM measurements. The authors acknowledge the Department of Chemistry of IUT at Castres for the facilities in synthesis, characterization and catalysis.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Damico, R. Preparation, Characterization, and Reactions of Lithium and Sodium Tetraalkylboron Compounds. J. Org. Chem. 1964, 29, 1971–1976. [Google Scholar] [CrossRef]
  2. Organic Syntheses, Inc. m-Chloroperbenzoic Acid. Org. Synth. 1970, 50, 15. [Google Scholar] [CrossRef]
  3. Brulé, E.; De Miguel, Y.R. Supported manganese porphyrin catalysts as P450 enzyme mimics for alkene epoxidation. Tetrahedron Lett. 2002, 43, 8555–8558. [Google Scholar] [CrossRef]
  4. Burfield, D.R.; Eng, A.-H. Glass transition and crystallization phenomena in epoxidized trans-polyisoprene: A differential scanning calorimetry study. Polymer 1989, 30, 2019–2022. [Google Scholar] [CrossRef]
  5. Dryuk, V.G. Advances in the Development of Methods for the Epoxidation of Olefins. Russ. Chem. Rev. 1985, 54, 986–1005. [Google Scholar] [CrossRef]
  6. Swern, D. (Ed.) Organic Peroxides; Interscience: New York, NY, USA, 1971; Volume 2. [Google Scholar]
  7. Shen, Y.; Jiang, P.; Wai, P.T.; Gu, Q.; Zhang, W. Recent Progress in Application of Molybdenum-Based Catalysts for Epoxidation of Alkenes. Catalysts 2019, 9, 31. [Google Scholar] [CrossRef] [Green Version]
  8. Srinivasan, K.; Michaud, P.; Kochi, J.K. Epoxidation of olefins with cationic (salen)manganese(III) complexes. The modulation of catalytic activity by substituents. J. Am. Chem. Soc. 1986, 108, 2309–2320. [Google Scholar] [CrossRef]
  9. Rudolph, J.; Reddy, K.L.; Chiang, J.P.; Sharpless, K.B. Highly Efficient Epoxidation of Olefins Using Aqueous H2O2 and Catalytic Methyltrioxorhenium/Pyridine: Pyridine-Mediated Ligand Acceleration. J. Am. Chem. Soc. 1997, 119, 6189–6190. [Google Scholar] [CrossRef]
  10. Kobayashi, M.; Tawara, K. Method for Producing Epoxy Compound. Japan Patent JP 2007230908A, 13 September 2007. [Google Scholar]
  11. Bagherzadeh, M.; Tahsini, L.; Latifi, R.; Woo, L.K. cis-Dioxo-molybdenum(VI)-oxazoline complex catalyzed epoxidation of olefins by tert-butyl hydrogen peroxide. Inorg. Chim. Acta 2009, 362, 3698–3702. [Google Scholar] [CrossRef]
  12. Dallmann, K.; Buffon, R.; Loh, W. Catalyst recycling in the epoxidation of alkenes catalyzed by MoO2(acac)2 through precipitation with poly(ethylene oxide). J. Mol. Catal. A Chem. 2002, 178, 43–46. [Google Scholar] [CrossRef]
  13. Yamazaki, M.; Endo, H.; Tomoyama, M.; Kurusu, Y. Catalytic Epoxidation of Cyclohexene witht-Butyl Hydroperoxide in the Presence of Various Molybdenum Complexes. Bull. Chem. Soc. Jpn. 1983, 56, 3523–3524. [Google Scholar] [CrossRef] [Green Version]
  14. Anderson, J.C.; Smith, N.M.; Robertson, M.; Scott, M.S. An investigation into oxo analogues of molybdenum olefin metathesis complexes as epoxidation catalysts for alkenes. Tetrahedron Lett. 2009, 50, 5344–5346. [Google Scholar] [CrossRef]
  15. Sherwood, J. European Restrictions on 1,2-Dichloroethane: C−H Activation Research and Development Should Be Liberated and not Limited. Angew. Chem. Int. Ed. 2018, 57, 14286–14290. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, W.; Agustin, D.; Poli, R. Influence of ligand substitution on molybdenum catalysts with tridentate Schiff base ligands for the organic solvent-free oxidation of limonene using aqueous TBHP as oxidant. Mol. Catal. 2017, 443, 52–59. [Google Scholar] [CrossRef]
  17. Wang, W.; Daran, J.-C.; Poli, R.; Agustin, D. OH-substituted tridentate ONO Schiff base ligands and related molybdenum(VI) complexes for solvent-free (ep)oxidation catalysis with TBHP as oxidant. J. Mol. Catal. A Chem. 2016, 416, 117–126. [Google Scholar] [CrossRef]
  18. Cvijanović, D.; Pisk, J.; Pavlović, G.; Šišak-Jung, D.; Matković-Čalogović, D.; Cindric, M.; Agustin, D.; Vrdoljak, V. Discrete mononuclear and dinuclear compounds containing a MoO22+ core and 4-aminobenzhydrazone ligands: Synthesis, structure and organic-solvent-free epoxidation activity. New J. Chem. 2018, 43, 1791–1802. [Google Scholar] [CrossRef]
  19. Cordelle, C.; Agustin, D.; Daran, J.-C.; Poli, R. Oxo-bridged bis oxo-vanadium(V) complexes with tridentate Schiff base ligands (VOL)2O (L=SAE, SAMP, SAP): Synthesis, structure and epoxidation catalysis under solvent-free conditions. Inorg. Chim. Acta 2010, 364, 144–149. [Google Scholar] [CrossRef]
  20. Morlot, J.; Uyttebroeck, N.; Agustin, D.; Poli, R. Solvent-Free Epoxidation of Olefins Catalyzed by “[MoO2(SAP)]”: A New Mode of tert -Butylhydroperoxide Activation. ChemCatChem 2012, 5, 601–611. [Google Scholar] [CrossRef]
  21. Guérin, B.; Fernandes, D.M.; Daran, J.-C.; Agustin, D.; Poli, R. Investigation of induction times, activity, selectivity, interface and mass transport in solvent-free epoxidation by H2O2 and TBHP: A study with organic salts of the [PMo12O40]3− anion. New J. Chem. 2013, 37, 3466–3475. [Google Scholar] [CrossRef]
  22. Pisk, J.; Agustin, D.; Poli, R. Organic Salts and Merrifield Resin Supported [PM12O40]3− (M = Mo or W) as Catalysts for Adipic Acid Synthesis. Molecules 2019, 24, 783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Wang, Y.; Gayet, F.; Guillo, P.; Agustin, D. Organic Solvent-Free Olefins and Alcohols (ep)oxidation Using Recoverable Catalysts Based on [PM12O40]3- (M = Mo or W) Ionically Grafted on Amino Functionalized Silica Nanobeads. Materials 2019, 12, 3278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Miao, C.; Wang, B.; Wang, Y.; Xia, C.; Lee, Y.-M.; Nam, W.; Sun, W. Proton-Promoted and Anion-Enhanced Epoxidation of Olefins by Hydrogen Peroxide in the Presence of Nonheme Manganese Catalysts. J. Am. Chem. Soc. 2016, 138, 936–943. [Google Scholar] [CrossRef] [PubMed]
  25. Ottenbacher, R.V.; Samsonenko, D.G.; Talsi, E.P.; Bryliakov, K.P. Enantioselective Epoxidations of Olefins with Various Oxidants on Bioinspired Mn Complexes: Evidence for Different Mechanisms and Chiral Additive Amplification. ACS Catal. 2016, 6, 979–988. [Google Scholar] [CrossRef]
  26. Du, J.; Miao, C.; Xia, C.; Lee, Y.-M.; Nam, W.; Sun, W. Mechanistic Insights into the Enantioselective Epoxidation of Olefins by Bioinspired Manganese Complexes: Role of Carboxylic Acid and Nature of Active Oxidant. ACS Catal. 2018, 8, 4528–4538. [Google Scholar] [CrossRef]
  27. Balleste, R.M.; Que, L. Iron-Catalyzed Olefin Epoxidation in the Presence of Acetic Acid: Insights into the Nature of the Metal-Based Oxidant. J. Am. Chem. Soc. 2007, 129, 15964–15972. [Google Scholar] [CrossRef] [PubMed]
  28. White, M.C.; Doyle, A.G.; Jacobsen, E.N. A synthetically useful, self-assembling MMO mimic system for catalytic alkene epoxidation with aqueous H2O2. J. Am. Chem. Soc. 2001, 123, 7194–7195. [Google Scholar] [CrossRef]
  29. Lorenz, S.; Plietker, B. Selectivity Trends in Olefin Epoxidations Catalyzed by (NNNN)Manganese(+II) Complexes using Trifluoroethanol as the Solvent. ChemCatChem 2016, 8, 3203–3206. [Google Scholar] [CrossRef]
  30. Duban, E.A.; Bryliakov, K.P.; Talsi, E.P. The Active Intermediates of Non-Heme-Iron-Based Systems for Catalytic Alkene Epoxidation with H2O2/CH3COOH. Eur. J. Inorg. Chem. 2007, 2007, 852–857. [Google Scholar] [CrossRef]
  31. Clemente-Tejeda, D.; Bermejo, F.A. Oxidation of alkenes with non-heme iron complexes: Suitability as an organic synthetic method. Tetrahedron 2014, 70, 9381–9386. [Google Scholar] [CrossRef]
  32. Clemente-Tejeda, D.; López-Moreno, A.; Bermejo, F.A. Non-heme iron catalysis in CC, C–H, and CH2 oxidation reactions. Oxidative transformations on terpenoids catalyzed by Fe(bpmen)(OTf)2. Tetrahedron 2013, 69, 2977–2986. [Google Scholar] [CrossRef]
  33. Clemente-Tejeda, D.; López-Moreno, A.; Bermejo, F.A. Oxidation of unsaturated steroid ketones with hydrogen peroxide catalyzed by Fe(bpmen)(OTf)2. New methodology to access biologically active steroids by chemo-, and stereoselective processes. Tetrahedron 2012, 68, 9249–9255. [Google Scholar] [CrossRef]
  34. Duban, E.A.; Brylyakov, K.P.; Talsi, E.P. The nature of active species in catalytic systems based on non-heme iron complexes, hydrogen peroxide, and acetic acid for selective olefin epoxidation. Kinet. Catal. 2008, 49, 379–385. [Google Scholar] [CrossRef]
  35. Taktak, S.; Kryatov, S.V.; Haas, T.E.; Rybak-Akimova, E.V. Diiron(III) oxo-bridged complexes with BPMEN and additional monodentate or bidentate ligands: Synthesis and reactivity in olefin epoxidation with H2O2. J. Mol. Catal. A Chem. 2006, 259, 24–34. [Google Scholar] [CrossRef]
  36. Chen, K.; Que, L., Jr. cis-Dihydroxylation of Olefins by a Non-Heme Iron Catalyst: A Functional Model for Rieske Dioxygenases. Angew. Chem. Int. Ed. 1999, 38, 2227–2229. [Google Scholar] [CrossRef]
  37. Cussó, O.; Garcia-Bosch, I.; Ribas, X.; Fillol, J.L.; Costas, M. Asymmetric Epoxidation with H2O2 by Manipulating the Electronic Properties of Non-heme Iron Catalysts. J. Am. Chem. Soc. 2013, 135, 14871–14878. [Google Scholar] [CrossRef] [PubMed]
  38. Bautz, J.; Comba, P.; De Laorden, C.L.; Menzel, M.; Rajaraman, G. Biomimetic High-Valent Non-Heme Iron Oxidants for thecis-Dihydroxylation and Epoxidation of Olefins. Angew. Chem. Int. Ed. 2007, 46, 8067–8070. [Google Scholar] [CrossRef] [PubMed]
  39. Yao, M.-Y.; Huang, Y.-B.; Niu, X.; Pan, H. Highly Efficient Silica-Supported Peroxycarboxylic Acid for the Epoxidation of Unsaturated Fatty Acid Methyl Esters and Vegetable Oils. ACS Sustain. Chem. Eng. 2016, 4, 3840–3849. [Google Scholar] [CrossRef]
  40. Crucho, C.I.C.; Baleizão, C.; Farinha, J.P.S. Functional Group Coverage and Conversion Quantification in Nanostructured Silica by 1H NMR. Anal. Chem. 2017, 89, 681–687. [Google Scholar] [CrossRef]
  41. Cohen, R.; Sukenik, C.N. Highly loaded COOH functionalized silica particles. Colloids Surf. A Physicochem. Eng. Asp. 2016, 504, 242–251. [Google Scholar] [CrossRef]
  42. Feinle, A.; Leichtfried, F.; Straßer, S.; Hüsing, N. Carboxylic acid-functionalized porous silica particles by a co-condensation approach. J. Sol-Gel Sci. Technol. 2017, 81, 138–146. [Google Scholar] [CrossRef] [Green Version]
  43. Boullanger, A.; Gracy, G.; Bibent, N.; Devautour-Vinot, S.; Clément, S.; Mehdi, A. From an Octakis(3-cyanopropyl)silsesquioxane Building Block to a Highly COOH-Functionalized Hybrid Organic-Inorganic Material. Eur. J. Inorg. Chem. 2011, 2012, 143–150. [Google Scholar] [CrossRef]
  44. Ghaida, F.A.; Clément, S.; Mehdi, A. Heterogenized Catalysis on Metals Impregnated Mesoporous Silica. In Novel Nanoscale Hybrid Materials; Wiley: Hoboken, NJ, USA, 2018; pp. 323–349. [Google Scholar]
  45. Touisni, N.; Kanfar, N.; Ulrich, S.; Dumy, P.; Supuran, C.T.; Mehdi, A.; Winum, J.-Y. Fluorescent Silica Nanoparticles with Multivalent Inhibitory Effects towards Carbonic Anhydrases. Chem.- Eur. J. 2015, 21, 10306–10309. [Google Scholar] [CrossRef]
  46. Chen, Y.; Zhou, Y.; Pi, H.; Zeng, G. Controlling the shear thickening behavior of suspensions by changing the surface properties of dispersed microspheres. RSC Adv. 2019, 9, 3469–3478. [Google Scholar] [CrossRef] [Green Version]
  47. Atta, S.; Fatima, M.; Islam, A.; Gull, N.; Sultan, M. Grafting of Silica Particles with Linoleic Acid via Modified Stober’s Method for Preconcenrtration of Pesticides in Drinking Water. Key Eng. Mater. 2018, 778, 316–324. [Google Scholar] [CrossRef] [Green Version]
  48. Yadav, M.; Akita, T.; Tsumori, N.; Xu, Q. Strong metal–molecular support interaction (SMMSI): Amine-functionalized gold nanoparticles encapsulated in silica nanospheres highly active for catalytic decomposition of formic acid. J. Mater. Chem. 2012, 22, 12582–12586. [Google Scholar] [CrossRef]
  49. Berg, R.V.D.; Parmentier, T.E.; Elkjær, C.F.; Gommes, C.; Sehested, J.; Helveg, S.; De Jongh, P.E.; De Jong, K.P. Support Functionalization To Retard Ostwald Ripening in Copper Methanol Synthesis Catalysts. ACS Catal. 2015, 5, 4439–4448. [Google Scholar] [CrossRef]
  50. Yantasee, W.; Rutledge, R.D.; Chouyyok, W.; Sukwarotwat, V.; Orr, G.; Warner, C.L.; Warner, M.G.; Fryxell, G.E.; Wiacek, R.J.; Timchalk, C.; et al. Functionalized Nanoporous Silica for the Removal of Heavy Metals from Biological Systems: Adsorption and Application. ACS Appl. Mater. Interfaces 2010, 2, 2749–2758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Kim, J.-S.; Chah, S.; Yi, J. Preparation of modified silica for heavy metal removal. Korean J. Chem. Eng. 2000, 17, 118–121. [Google Scholar] [CrossRef]
  52. Leon, P.A.A.I.-D.; Contreras, C.A.; Thornburg, N.E.; Thompson, A.B.; Notestein, J.M. Catalyst structure and substituent effects on epoxidation of styrenics with immobilized Mn(tmtacn) complexes. Appl. Catal. A Gen. 2016, 511, 78–86. [Google Scholar] [CrossRef]
  53. Schoenfeldt, N.J.; Ni, Z.; Korinda, A.W.; Meyer, R.J.; Notestein, J.M. Manganese Triazacyclononane Oxidation Catalysts Grafted under Reaction Conditions on Solid Cocatalytic Supports. J. Am. Chem. Soc. 2011, 133, 18684–18695. [Google Scholar] [CrossRef]
  54. Ottenbacher, R.V.; Talsi, E.P.; Bryliakov, K.P. Bioinspired Mn-aminopyridine catalyzed epoxidations of olefins with various oxidants: Enantioselectivity and mechanism. Catal. Today 2016, 278, 30–39. [Google Scholar] [CrossRef]
  55. Cussó, O.; Serrano-Plana, J.; Costas, M. Evidence of a Sole Oxygen Atom Transfer Agent in Asymmetric Epoxidations with Fe-pdp Catalysts. ACS Catal. 2017, 7, 5046–5053. [Google Scholar] [CrossRef]
  56. Hureau, C.; Blondin, G.; Charlot, M.-F.; Philouze, C.; Nierlich, M.; Cesario, M.; Anxolabéhère-Mallart, E. Synthesis, Structure, and Characterization of New Mononuclear Mn(II) Complexes. Electrochemical Conversion into New Oxo-Bridged Mn2(III,IV) Complexes. Role of Chloride Ions. Inorg. Chem. 2005, 44, 3669–3683. [Google Scholar] [CrossRef]
  57. Chow, T.W.-S.; Wong, E.L.-M.; Guo, Z.; Liu, Y.; Huang, J.-S.; Che, C.M. cis-Dihydroxylation of Alkenes with Oxone Catalyzed by Iron Complexes of a Macrocyclic Tetraaza Ligand and Reaction Mechanism by ESI-MS Spectrometry and DFT Calculations. J. Am. Chem. Soc. 2010, 132, 13229–13239. [Google Scholar] [CrossRef]
  58. To, W.-P.; Chow, T.W.-S.; Tse, C.-W.; Guan, X.; Huang, J.-S.; Che, C.-M. Water oxidation catalysed by iron complex of N,N′-dimethyl-2,11-diaza[3,3](2,6)pyridinophane. Spectroscopy of iron–oxo intermediates and density functional theory calculations. Chem. Sci. 2015, 6, 5891–5903. [Google Scholar] [CrossRef] [Green Version]
  59. Murphy, A.; Dubois, G.; Stack, T.D.P. Efficient Epoxidation of Electron-Deficient Olefins with a Cationic Manganese Complex. J. Am. Chem. Soc. 2003, 125, 5250–5251. [Google Scholar] [CrossRef] [PubMed]
  60. Dexuan, W.; Guian, L.; Qingyan, H.; Ziqiang, W.; Liping, P.; Zhongyue, Z.; Hairong, Z. Synthesis of Au-SiO2 Composite Nanospheres and Their Catalytic Activity. J. Nanosci. Nanotechnol. 2016, 16, 3821–3826. [Google Scholar] [CrossRef]
  61. Bourebrab, M.A.; Oben, D.T.; Durand, G.G.; Taylor, P.G.; Bruce, J.I.; Bassindale, A.R.; Taylor, A. Influence of the initial chemical conditions on the rational design of silica particles. J. Sol-Gel Sci. Technol. 2018, 88, 430–441. [Google Scholar] [CrossRef] [Green Version]
  62. Green, D.; Jayasundara, S.; Lam, Y.-F.; Harris, M. Chemical reaction kinetics leading to the first Stober silica nanoparticles—NMR and SAXS investigation. J. Non-Cryst. Solids 2003, 315, 166–179. [Google Scholar] [CrossRef]
  63. Suratwala, T.; Hanna, M.; Whitman, P. Effect of humidity during the coating of Stöber silica sols. J. Non- Cryst. Solids 2004, 349, 368–376. [Google Scholar] [CrossRef]
  64. Stöber, W.; Fink, A.; Bohn, E. Controlled growth of monodisperse silica spheres in the micron size range. J. Colloid Interface Sci. 1968, 26, 62–69. [Google Scholar] [CrossRef]
  65. Wang, X.-D.; Shen, Z.-X.; Sang, T.; Cheng, X.-B.; Li, M.-F.; Chen, L.-Y.; Wang, Z.-S. Preparation of spherical silica particles by Stöber process with high concentration of tetra-ethyl-orthosilicate. J. Colloid Interface Sci. 2010, 341, 23–29. [Google Scholar] [CrossRef]
  66. Green, D.; Lin, J.; Lam, Y.-F.; Hu, M.; Schaefer, D.W.; Harris, M. Size, volume fraction, and nucleation of Stober silica nanoparticles. J. Colloid Interface Sci. 2003, 266, 346–358. [Google Scholar] [CrossRef]
  67. Malay, O.; Yilgor, I.; Menceloglu, Y.Z. Effects of solvent on TEOS hydrolysis kinetics and silica particle size under basic conditions. J. Sol-Gel Sci. Technol. 2013, 67, 351–361. [Google Scholar] [CrossRef]
  68. Bu, J.; Li, R.; Quah, C.W.; Carpenter, K.J. Propagation of PAMAM Dendrons on Silica Gel: A Study on the Reaction Kinetics. Macromolecules 2004, 37, 6687–6694. [Google Scholar] [CrossRef]
  69. Aneja, K.S.; Bohm, S.; Khanna, A.S.; Bohm, H.L.M. Graphene based anticorrosive coatings for Cr(vi) replacement. Nanoscale 2015, 7, 17879–17888. [Google Scholar] [CrossRef]
  70. Das, D.; Yang, Y.; O’Brien, J.S.; Breznan, D.; Nimesh, S.; Bernatchez, S.; Hill, M.; Sayari, A.; Vincent, R.; Kumarathasan, P. Synthesis and Physicochemical Characterization of MesoporousSiO2Nanoparticles. J. Nanomater. 2014, 2014, 1–12. [Google Scholar] [CrossRef] [Green Version]
  71. Feifel, S.C.; Lisdat, F. Silica nanoparticles for the layer-by-layer assembly of fully electro-active cytochrome c multilayers. J. Nanobiotechnology 2011, 9, 59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Trébosc, J.; Wiench, J.W.; Huh, S.; Lin, V.S.-Y.; Pruski, M. Solid-State NMR Study of MCM-41-type Mesoporous Silica Nanoparticles. J. Am. Chem. Soc. 2005, 127, 3057–3068. [Google Scholar] [CrossRef] [PubMed]
  73. Mouawia, R.; Mehdi, A.; Reyé, C.; Corriu, R. Direct synthesis of ordered and highly functionalized organosilicas containing carboxylic acid groups. J. Mater. Chem. 2007, 17, 616–618. [Google Scholar] [CrossRef]
  74. Sharma, R.K.; Sharma, S.; Gulati, S.; Pandey, A. Fabrication of a novel nano-composite carbon paste sensor based on silica-nanospheres functionalized with isatin thiosemicarbazone for potentiometric monitoring of Cu2+ ions in real samples. Anal. Methods 2013, 5, 1414–1426. [Google Scholar] [CrossRef]
  75. Ribeiro, S.O.; Granadeiro, C.; de Almeida, P.M.; Pires, J.; Sánchez, M.D.C.C.; Campos-Martin, J.M.; Gago, S.; de Castro, B.; Balula, S.S. Oxidative desulfurization strategies using Keggin-type polyoxometalate catalysts: Biphasic versus solvent-free systems. Catal. Today 2019, 333, 226–236. [Google Scholar] [CrossRef]
  76. Park, J.S.; Hah, J.; Koo, S.M.; Lee, Y.S. Effect of Alcohol Chain Length on Particle Growth in a Mixed Solvent System. J. Ceram. Process. Res. 2006, 7, 83–89. [Google Scholar]
  77. Beganskienė, A.; Sirutkaitis, V.; Kurtinaitienė, M.; Juškėnas, R.; Kareiva, A. FTIR, TEM and NMR Iinvestigations of Stöber Silica Nanoparticles. Mater. Sci. (Medžiagotyra) 2004, 10, 287–290. [Google Scholar] [CrossRef]
  78. Van De Vyver, S.; Roman-Leshkov, Y. Emerging catalytic processes for the production of adipic acid. Catal. Sci. Technol. 2013, 3, 1465–1479. [Google Scholar] [CrossRef] [Green Version]
  79. Cavani, F.; Alini, S. Synthesis of Adipic Acid: On the Way to More Sustainable Production. In Sustainable Industrial Processes; Cavani, F., Centi, G., Perathoner, S., Trifiro, F., Eds.; Wiley: Weinheim, Germany, 2009; pp. 367–426. [Google Scholar]
  80. Chen, K.; Costas, M.; Kim, J.; Tipton, A.A.K.; Que, J.L. Olefin Cis-Dihydroxylation versus Epoxidation by Non-Heme Iron Catalysts: Two Faces of an FeIII−OOH Coin. J. Am. Chem. Soc. 2002, 124, 3026–3035. [Google Scholar] [CrossRef]
  81. Gelasco, A.; Askenas, A.; Pecoraro, V.L. Catalytic Disproportionation of Hydrogen Peroxide by the Tetranuclear Manganese Complex [Mnll(2-OHpicpn)]4. Inorg. Chem. 1996, 35, 1419–1420. [Google Scholar] [CrossRef]
  82. Fenton, H.J.H. LXXIII.—Oxidation of tartaric acid in presence of iron. J. Chem. Soc. Trans. 1894, 65, 899–910. [Google Scholar] [CrossRef] [Green Version]
  83. Jaouen, F.; Dodelet, J.-P. O2 Reduction Mechanism on Non-Noble Metal Catalysts for PEM Fuel Cells. Part I: Experimental Rates of O2 Electroreduction, H2O2 Electroreduction, and H2O2 Disproportionation. J. Phys. Chem. C 2009, 113, 15422–15432. [Google Scholar] [CrossRef]
  84. Sengupta, K.; Chatterjee, S.; Dey, A. Catalytic H2O2 Disproportionation and Electrocatalytic O2 Reduction by a Functional Mimic of Heme Catalase: Direct Observation of Compound 0 and Compound I in Situ. ACS Catal. 2016, 6, 1382–1388. [Google Scholar] [CrossRef] [Green Version]
  85. Nourian, M.; Zadehahmadi, F.; Kardanpour, R.; Tangestaninejad, S.; Moghadam, M.; Mirkhani, V.; Mohammadpoor-Baltork, I. Highly efficient oxidative cleavage of alkenes and cyanosilylation of aldehydes catalysed by magnetically recoverable MIL-101. Appl. Organomet. Chem. 2018, 32, e3957. [Google Scholar] [CrossRef]
  86. Wang, J.Y.; Zhou, M.D.; Yuan, Y.G.; Fu, N.H.; Zang, S.L. Oxidation of cyclooctene to suberic acid using perrhenate-containing composite ionic liquids as green catalysts. Russ. J. Gen. Chem. 2015, 85, 2378–2385. [Google Scholar] [CrossRef]
  87. Chen, J.; Chen, M.; Zhang, B.; Nie, R.; Huang, A.; Goh, T.W.; Volkov, A.; Zhang, Z.; Ren, Q.; Huang, W. Allylic oxidation of olefins with a manganese-based metal–organic framework. Green Chem. 2019, 21, 3629–3636. [Google Scholar] [CrossRef]
  88. Chavan, S.; Srinivas, D.; Ratnasamy, P. Oxidation of Cyclohexane, Cyclohexanone, and Cyclohexanol to Adipic Acid by a Non-HNO3 Route over Co/Mn Cluster Complexes. J. Catal. 2002, 212, 39–45. [Google Scholar] [CrossRef]
  89. Schuchardt, U.; Cardoso, D.; Sercheli, R.; Pereira, R.; da Cruz, R.S.; Guerreiro, M.C.; Mandelli, D.; Spinacé, E.V.; Pires, E.L. Cyclohexane oxidation continues to be a challenge. Appl. Catal. A Gen. 2001, 211, 1–17. [Google Scholar] [CrossRef]
  90. Shen, D.; Miao, C.; Xu, D.; Xia, C.; Sun, W. Highly Efficient Oxidation of Secondary Alcohols to Ketones Catalyzed by Manganese Complexes of N4 Ligands with H2O2. Org. Lett. 2014, 17, 54–57. [Google Scholar] [CrossRef] [PubMed]
  91. Nehru, K.; Kim, S.J.; Kim, I.Y.; Seo, M.S.; Kim, Y.; Kim, S.-J.; Kim, J.; Nam, W. A highly efficient non-heme manganese complex in oxygenation reactions. Chem. Commun. 2007, 44, 4623–4625. [Google Scholar] [CrossRef]
  92. Andraos, J.; Sayed, M. On the Use of “Green” Metrics in the Undergraduate Organic Chemistry Lecture and Lab To Assess the Mass Efficiency of Organic Reactions. J. Chem. Educ. 2007, 84, 1004. [Google Scholar] [CrossRef]
  93. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  94. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A Found. Crystallogr. 2008, 64, 112–122. [Google Scholar] [CrossRef] [Green Version]
  95. Farrugia, L.J. ORTEP-3 for Windows—A version ofORTEP-III with a Graphical User Interface (GUI). J. Appl. Crystallogr. 1997, 30, 565. [Google Scholar] [CrossRef]
  96. Burnett, M.N.; Johnson, C.K. ORTEPIII. Report ORNL-6895; Oak Ridge National Laboratory: Oak Ridge, TN, USA, 1996.
Figure 1. Synthesis of metal complexes of L.
Figure 1. Synthesis of metal complexes of L.
Molecules 26 05435 g001
Figure 2. Molecular views of (L)Mn(p-Ts)2 (a) and [(L)FeCl2](FeCl4) (b) with the atom labelling scheme. Ellipsoids are drawn at the 50% probability level. H atoms have been omitted for the sake of clarity for (L)Mn(p-Ts)2.
Figure 2. Molecular views of (L)Mn(p-Ts)2 (a) and [(L)FeCl2](FeCl4) (b) with the atom labelling scheme. Ellipsoids are drawn at the 50% probability level. H atoms have been omitted for the sake of clarity for (L)Mn(p-Ts)2.
Molecules 26 05435 g002
Figure 3. Synthesis of SiO2 particles.
Figure 3. Synthesis of SiO2 particles.
Molecules 26 05435 g003
Figure 4. Synthetic pathway of the functionalized SiO2 nanoparticles.
Figure 4. Synthetic pathway of the functionalized SiO2 nanoparticles.
Molecules 26 05435 g004
Figure 5. From top to bottom: TEM images and diameter distribution of SiO2, SiO2@CN, SiO2@COOH beads from SiO2 beads produced in EtOH (a) and MeOH (b).
Figure 5. From top to bottom: TEM images and diameter distribution of SiO2, SiO2@CN, SiO2@COOH beads from SiO2 beads produced in EtOH (a) and MeOH (b).
Molecules 26 05435 g005
Figure 6. From top to bottom: size (hydrodynamic radius) distribution (in number) obtained by DLS for SiO2, SiO2@CN, SiO2@COOH beads from SiO2 beads produced in EtOH (a) and MeOH (b).
Figure 6. From top to bottom: size (hydrodynamic radius) distribution (in number) obtained by DLS for SiO2, SiO2@CN, SiO2@COOH beads from SiO2 beads produced in EtOH (a) and MeOH (b).
Molecules 26 05435 g006
Figure 7. Relevant IR vibration zones for SiO2 (a), SiO2@CN (b), SiO2@COOH (c) beads from SiO2 beads produced in EtOH (A) and MeOH (B).
Figure 7. Relevant IR vibration zones for SiO2 (a), SiO2@CN (b), SiO2@COOH (c) beads from SiO2 beads produced in EtOH (A) and MeOH (B).
Molecules 26 05435 g007
Figure 8. Difference spectra (SiO2@CN-SiO2) on two specific ranges, i.e 400-1500 cm−1 (A) and 2200-2500 cm−1 (B). The spectrum of TESPN is indicated in (a), (b) SiO2 produced in MeOH, (c) with SiO2 produced in EtOH.
Figure 8. Difference spectra (SiO2@CN-SiO2) on two specific ranges, i.e 400-1500 cm−1 (A) and 2200-2500 cm−1 (B). The spectrum of TESPN is indicated in (a), (b) SiO2 produced in MeOH, (c) with SiO2 produced in EtOH.
Molecules 26 05435 g008
Figure 9. Difference spectra (SiO2@COOH-SiO2) on specific range. (a) with SiO2 produced in MeOH, (b) with SiO2 produced in EtOH.
Figure 9. Difference spectra (SiO2@COOH-SiO2) on specific range. (a) with SiO2 produced in MeOH, (b) with SiO2 produced in EtOH.
Molecules 26 05435 g009
Figure 10. 29Si CPMAS NMR spectra of SiO2 (a) SiO2@CN (b), SiO2@COOH (c) from SiO2 produced in EtOH (A) and MeOH (B).
Figure 10. 29Si CPMAS NMR spectra of SiO2 (a) SiO2@CN (b), SiO2@COOH (c) from SiO2 produced in EtOH (A) and MeOH (B).
Molecules 26 05435 g010
Figure 11. Schematic functions on the silica beads.
Figure 11. Schematic functions on the silica beads.
Molecules 26 05435 g011
Figure 12. Schematic representation of the silica beads
Figure 12. Schematic representation of the silica beads
Molecules 26 05435 g012
Figure 13. Catalytic oxidation of cyclooctene.
Figure 13. Catalytic oxidation of cyclooctene.
Molecules 26 05435 g013
Figure 14. Comparison of CO conversion between different conditions for (L)MnCl2 (a), (L)Mn(OTf)2 (b), (L)Mn(p-Ts)2 (c), (L)FeCl2(FeCl4) (d).
Figure 14. Comparison of CO conversion between different conditions for (L)MnCl2 (a), (L)Mn(OTf)2 (b), (L)Mn(p-Ts)2 (c), (L)FeCl2(FeCl4) (d).
Molecules 26 05435 g014
Figure 15. Catalytic oxidation of cyclohexene.
Figure 15. Catalytic oxidation of cyclohexene.
Molecules 26 05435 g015
Figure 16. Comparison of conversion (%) of CH between different catalysts (L)MnCl2 (a), (L)Mn(OTf)2 (b), (L)Mn(p-Ts)2 (c), (L)FeCl2(FeCl4) (d) and different co-reagents. Reaction time: 3 h with CH3COOH, 5 h with SiO2@COOH.
Figure 16. Comparison of conversion (%) of CH between different catalysts (L)MnCl2 (a), (L)Mn(OTf)2 (b), (L)Mn(p-Ts)2 (c), (L)FeCl2(FeCl4) (d) and different co-reagents. Reaction time: 3 h with CH3COOH, 5 h with SiO2@COOH.
Molecules 26 05435 g016
Figure 17. Catalytic oxidation of cyclohexanol.
Figure 17. Catalytic oxidation of cyclohexanol.
Molecules 26 05435 g017
Figure 18. Comparison of CYol conversion (%) between different catalysts (L)MnCl2 (a), (L)Mn(OTf)2 (b), (L)Mn(p-Ts)2 (c), (L)FeCl2(FeCl4) (d) and different co-reagents. Reaction time: 3 h with CH3COOH, 5 h with SiO2@COOH.
Figure 18. Comparison of CYol conversion (%) between different catalysts (L)MnCl2 (a), (L)Mn(OTf)2 (b), (L)Mn(p-Ts)2 (c), (L)FeCl2(FeCl4) (d) and different co-reagents. Reaction time: 3 h with CH3COOH, 5 h with SiO2@COOH.
Molecules 26 05435 g018
Figure 19. Comparison of green metrics for the epoxidation of cyclooctene (yield, RME and MRP) with the different catalysts and the different co-reagents acetic acid (blue), SiO2@COOH(E) (orange) and SiO2@COOH(M) (grey).
Figure 19. Comparison of green metrics for the epoxidation of cyclooctene (yield, RME and MRP) with the different catalysts and the different co-reagents acetic acid (blue), SiO2@COOH(E) (orange) and SiO2@COOH(M) (grey).
Molecules 26 05435 g019
Figure 20. Comparison of green metrics for the epoxidation of cyclohexene (the yield considered the cyclohexene oxide only) (yield, RME and MRP) with the different catalysts and the different co-reagents acetic acid (blue), SiO2@COOH(E) (orange) and SiO2@COOH(M) (grey).
Figure 20. Comparison of green metrics for the epoxidation of cyclohexene (the yield considered the cyclohexene oxide only) (yield, RME and MRP) with the different catalysts and the different co-reagents acetic acid (blue), SiO2@COOH(E) (orange) and SiO2@COOH(M) (grey).
Molecules 26 05435 g020
Figure 21. Comparison of green metrics for the oxidation of cyclohexanol (yield, RME and MRP) with the different catalysts and the different co-reagents acetic acid (blue), SiO2@COOH(E) (orange) and SiO2@COOH(M) (grey).
Figure 21. Comparison of green metrics for the oxidation of cyclohexanol (yield, RME and MRP) with the different catalysts and the different co-reagents acetic acid (blue), SiO2@COOH(E) (orange) and SiO2@COOH(M) (grey).
Molecules 26 05435 g021
Table 1. Selected bond distances (Å) and angles (deg.) for (L)Mn(p-Ts)2 and [(L)FeCl2](FeCl4).
Table 1. Selected bond distances (Å) and angles (deg.) for (L)Mn(p-Ts)2 and [(L)FeCl2](FeCl4).
(L)Mn(p-Ts)2[(L)FeCl2](FeCl4)
Bonds (Å)
M-Npy2.308(5)–2.352(2)2.1408(12), 2.1556(12)
M-Namine2.249(2)–2.283(2)2.2233(11), 2.2264(12)
Angles (°)
Namine-M-Namine75.46(9)–76.06(8)79.70(4)
NPy-M-NPy168.55(8)–168.87(7)166.17(5)
Table 2. Number of functions (F) (mmol) per g sample, calculated by 1H NMR.
Table 2. Number of functions (F) (mmol) per g sample, calculated by 1H NMR.
Sρ(f) (mmol F/g S)
OCH2CH3OCH3CNCOOH
SiO2 (E)0.43
SiO2@CN (E)0.64 0.29
SiO2@COOH (E)0.45 0.04
SiO2 (M)1.180.05
SiO2@CN (M)1.850.041.40
SiO2@COOH (M)0.080.05 0.31
Table 3. Number of function (mol) per nm2 core (μ(f)).
Table 3. Number of function (mol) per nm2 core (μ(f)).
Solvent Used
for SiO2 Synthesis
SiO2@CNSiO2@COOH
Ethanol20.62.8
Methanol16.63.2
Table 4. Relevant data for the catalyzed epoxidation of CO.(a).
Table 4. Relevant data for the catalyzed epoxidation of CO.(a).
CatalystRCOOHCOCOETON (e)
Conv (b)Sel (c)Yield (d)
(L)MnCl2no1---
CH3COOH998181100
CH3COOH (f)1---
SiO2@COOH(M)379438
SiO2@COOH(E)55261455
(L)Mn(OTf)2no57<13
CH3COOH99545499
SiO2@COOH(M)50452350
SiO2@COOH(E)53432352
(L)Mn(p-Ts)2no5502.76
CH3COOH1006262100
SiO2@COOH(M)61301961
SiO2@COOH(E)62282362
[(L)FeCl2](FeCl4)no0---
CH3COOH60211360
SiO2@COOH(M)80312580
SiO2@COOH(E)91252391
(a) Experimental conditions: 0 °C with CH3COOH, 60 °C with SiO2@COOH. Cat/H2O2/CO/CH3COOH = 1/150/100/1400 for CH3COOH, t = 3 h; Cat/H2O2/CO/COOH = 1/150/100/14 for SiO2@COOH, t = 5 h. (b) nCO converted/nCO engaged (%) at the end of the reaction. (c) nCOE formed/nCO converted at the end of the reaction. (d) nCOE formed/nCO engaged at the end of the reaction. (e) nCO transformed/ncat at the end of the reaction. (f) Cat/H2O2/CO/CH3COOH=1/150/100/14, t = 3 h, 0 °C.
Table 5. Relevant data for the catalyzed (ep)oxidation of cyclohexene (a).
Table 5. Relevant data for the catalyzed (ep)oxidation of cyclohexene (a).
CatalystRCOOHConv (b)Selectivity (c)TON (d)
CHCHOCHDCHolCHone
(L)MnCl2CH3COOH10089000100
SiO2@COOH(M)633.302263
SiO2@COOH(E)7414230074
(L)Mn(OTf)2CH3COOH985730198
SiO2@COOH(M)571300056
SiO2@COOH(E)832700083
(L)Mn(p-Ts)2CH3COOH966820296
SiO2@COOH(M)531600053
SiO2@COOH(E)802800080
[(L)FeCl2](FeCl4)CH3COOH11000011
SiO2@COOH(M)8792361786
SiO2@COOH(E)96450996
(a) Conditions: 0 °C for the case with CH3COOH, 60 °C for the case with SiO2@COOH. Cat/H2O2/CH/CH3COOH = 1/150/100/1400 for CH3COOH, t = 3 h; Cat/H2O2/CH/COOH = 1/150/100/14 for SiO2@COOH, t = 5 h. (b) nCH converted/nCH engaged (in%) after 3 h for CH3COOH, 5 h for SiO2@COOH. (c) nproduct formed/ nCH converted at 3 h for CH3COOH, 5 h for SiO2@COOH. (d) nCH transformed /nCat at 3 h for CH3COOH, 5 h for SiO2@COOH.
Table 6. Relevant data for the catalyzed oxidation of cyclohexanol (a).
Table 6. Relevant data for the catalyzed oxidation of cyclohexanol (a).
CatalystRCOOHCYolCYoneTON (e)
Conv (b)Sel (c)Yield (d)
(L)MnCl2CH3COOH81917481
SiO2@COOH(M)1546715
SiO2@COOH(E)16901416
(L)Mn(OTf)2CH3COOH1007979100
SiO2@COOH(M)23902123
SiO2@COOH(E)27872427
(L)Mn(p-Ts)2CH3COOH99858599
SiO2@COOH(M)21972121
SiO2@COOH(E)25872225
[(L)FeCl2](FeCl4)CH3COOH100797999
SiO2@COOH(M)59452759
SiO2@COOH(E)65362365
(a) Conditions: 0 °C for the case with CH3COOH, 60 °C for the case with SiO2@COOH Cat/H2O2/CYol/CH3COOH = 1/150/100/1400 for CH3COOH, t = 3 h; Cat/H2O2/CYol/COOH = 1/150/100/14 for SiO2@COOH, t = 5 h. (b) nCYol converted/nCYol engaged (in%) after 3 h for CH3COOH, 5 h for SiO2@COOH. (c) nCYone formed/ nCYol converted at 3 h for CH3COOH, 5h for SiO2@COOH. (d) nCYone formed/ nCYol engaged at 3h for CH3COOH, 5 h for SiO2@COOH. (e) nCYol transformed /nCat at 3 h for CH3COOH, 5 h for SiO2@COOH.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, Y.; Gayet, F.; Daran, J.-C.; Guillo, P.; Agustin, D. Replacement of Volatile Acetic Acid by Solid SiO2@COOH Silica (Nano)Beads for (Ep)Oxidation Using Mn and Fe Complexes Containing BPMEN Ligand. Molecules 2021, 26, 5435. https://doi.org/10.3390/molecules26185435

AMA Style

Wang Y, Gayet F, Daran J-C, Guillo P, Agustin D. Replacement of Volatile Acetic Acid by Solid SiO2@COOH Silica (Nano)Beads for (Ep)Oxidation Using Mn and Fe Complexes Containing BPMEN Ligand. Molecules. 2021; 26(18):5435. https://doi.org/10.3390/molecules26185435

Chicago/Turabian Style

Wang, Yun, Florence Gayet, Jean-Claude Daran, Pascal Guillo, and Dominique Agustin. 2021. "Replacement of Volatile Acetic Acid by Solid SiO2@COOH Silica (Nano)Beads for (Ep)Oxidation Using Mn and Fe Complexes Containing BPMEN Ligand" Molecules 26, no. 18: 5435. https://doi.org/10.3390/molecules26185435

Article Metrics

Back to TopTop