Next Article in Journal
On the Exceptionally High Loading of L-Proline on Multi-Wall Carbon Nanotubes
Next Article in Special Issue
Microporous Zeolites and Related Nanoporous Materials: Synthesis, Characterization and Application in Catalysis
Previous Article in Journal
Catalysts Prepared with Matured Compost Derived from Mechanical-Biological Treatment Plants for the Wet Peroxide Oxidation of Pollutants with Different Lipophilicity
Previous Article in Special Issue
CHA-Type Zeolite Prepared by Interzeolite Conversion Method Using FAU and LTL-Type Zeolite: Effect of the Raw Materials on the Crystallization Mechanism, and Physicochemical and Catalytic Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Solvent Influence on Selectivity in α-Pinene Oxide Isomerization Using MoO3-Modified Zeolite BETA

1
Department of Organic Technology, University of Chemistry and Technology, 16628 Prague, Czech Republic
2
Department of Inorganic Technology, University of Chemistry and Technology, 16628 Prague, Czech Republic
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(11), 1244; https://doi.org/10.3390/catal10111244
Submission received: 28 September 2020 / Revised: 22 October 2020 / Accepted: 23 October 2020 / Published: 28 October 2020

Abstract

:
Natural source turpentine is an available source of α-pinene oxide. This compound’s value is especially given by the possibility of producing important compounds campholenic aldehyde and trans-carveol. In this work, we would like to present the usage of MoO3-modified zeolite BETA in α-pinene oxide isomerization concerning campholenic aldehyde and trans-carveol formation using a wide range of solvents. Catalyst calcination temperature also influenced the reaction course (selectivity to desired compounds and reaction rate). MoO3-zeolite BETA was prepared by the wet impregnation method and characterized by different techniques. The use of polar aprotic solvents had the most positive effect on the reaction course. Solvent basicity and polarity considerably influenced the reaction rate and selectivity to particular products. The combination of high basicity and the high polarity was the most suitable for the studied reaction from the reaction rate point of view. Selectivity to campholenic aldehyde and trans-carveol was the most influenced by solvent basicity. Higher solvent basicity caused the preferential formation of trans–carveol, influence on selectivity to campholenic aldehyde formation was the opposite. The described catalyst may be used for α-pinene oxide rearrangement to both desired products dependently on the used solvent. Molybdenum offers an exciting alternative for previously described modifications of zeolites for this reaction.

Graphical Abstract

1. Introduction

α-Pinene oxide is an important compound, which does not occur in nature itself. Still, simple oxidation of α-pinene—a natural compound abundant present in renewable source turpentine (obtained from biomass processing) [1]—leads to α-pinene oxide. The most valued compounds, which can be obtained from this compound isomerization, are campholenic aldehyde and trans-carveol. Campholenic aldehyde, which itself possesses by a gently woody smell, can be further used to synthesize fragrances with sandalwood scent (e.g., sandalore, Figure 1), trans-carveol can be part of fragrances bearing caraway and spearmint scent.
Several biological activities (antimicrobial, anticancer [2,3]) accompany these substances. Abrushov synthesis [4,5] serves as a process for campholenic aldehyde preparation using homogeneous catalyst zinc (II) bromide biotransformation of d-limonene led to trans-carveol [6]. Usage of heterogeneous catalysts in this reaction can offer environmental aspects (not handling corrosive mixtures, the possibility of catalyst reuse). Many researchers gave attention to the study of various heterogeneous catalysts in α-pinene oxide isomerization in recent years. Many catalysts based on (alumo)silicates was already tested, among them iron-modified zeolites or mesoporous sieves, providing > 60% selectivity of campholenic aldehyde formation [7,8,9,10,11,12,13]. Preferential formation of trans-carveol [14,15] was described using titanium modified materials [16,17,18], Sn-SiO2 with 70% selectivity of trans-carveol formation [19], methyl(trioxo)rhenium-Al2O3 providing total conversion in a single minute [20], and 83% selectivity of trans-carveol formation, Ce-MCM-41 [21] or tungsten modified SiO2 [22,23]. Some literary data describe the influence of solvent on selectivity in α-pinene oxide rearrangement toward particular products. Above all, using iron- or titanium-modified materials [7,9,14,16,17]—overall results showed that usage of basic polar solvents favored in most cases the formation of trans-carveol and using of nonpolar solvents campholenic aldehyde formation. On the other hand, selectivity which product occurs from α-pinene oxide rearrangement is given by a specific combination of temperature, catalyst, and solvent type (e.g., modified clays—high selectivity of fencholenic aldehyde formation was reported [24]).
Molybdenum modified materials are active acid catalysts (because of material Lewis acidity enhancement due MoO3 content [25,26]) and may serve as catalysts in alkane isomerization [27,28,29,30,31]. Usage of molybdenum-based catalyst in α-pinene oxide is also reported [32]. Based on these facts, we decided to prepare several types of molybdenum modified zeolites BETA and test a large number of different solvents concerning campholenic aldehyde and trans-carveol selectivity.

2. Results and Discussion

2.1. Material Preparation and Characterization

Several types of Mo-modified materials were prepared by wet impregnation method of BETA38 zeolite. The wet impregnation method is commonly used for zeolite modification and was performed using two temperatures based on the literature data: 60 °C [10,14] (for materials 20Mo450, 20Mo500, 20Mo550, 20Mo600) and RT (20Mo450RT) [25,26]. The effect of calcination temperature on the material properties and catalytic activity was also tested—calcination temperatures 450, 500, 550 and 600 °C (used calcination temperature is contained in materials name—e.g., 20Mo550 was calcined at 550 °C).
To determine the optimal calcination temperature, the temperature programmed heating was applied to non-calcined material precursor obtained after the wet impregnation method (Figure 2). TCD (thermal conductivity detector) signal helped us to determine the temperature at which all molybdate decomposed to MoO3. This temperature was about 450 °C. We applied four calcination temperatures starting from 450°C on material precursor in connection with this finding.
X-ray fluorescence (XRF) analysis of samples confirmed that all samples contained appropriate Mo amount corresponding to the desired (20 wt.%) content of molybdenum in catalyst (Table 1).
X-ray diffraction (Figure 3) confirmed the crystal structure of zeolite BETA38—characteristic bands for alumosilicate (7.8, 12.8, 14.6, 21.4, 22.5, 25.4, 25.6, 27.3, 28.9, 29.6, 43.6 2Theta°) were present [33]. Materials modified with MoO3 contained aforementioned characteristic bands (with the exception of 20Mo600) for alumosilicate. This fact revealed that inserting MoO3 did not destroy the zeolite structure. Bands characteristic for MoO3 (12.8, 23.3, 27.3, 33.2, 33.8, 38.9, 39.5, 45.8, 46.3, 49.3, 52.8, 55.2, 56.4, 58.8, 64.5, 69.6, 72.8, 76.5, 78.8 2Theta°) [34] were also present. The presence of intensive bands for MoO3 corresponded to the fact that the sample contained a large amount of MoO3, which was not homogenously dispersed [35]. Materials 20Mo550 and 20Mo600 were partially amorphous, which was especially visible by the absence of a band at 7.8 2Theta° and a decrease of some other bands’ intensity characteristic for alumosilicate.
Table 2 summarizes the results of the textural analyses of the catalysts. All samples of catalysts show adsorption isotherms of type Ib and IVa (Figure 4, IUPAC classification), so samples consist of micropores and mesopores with classical type hysteresis loop H1 (IUPAC classification), which corresponds to a narrow range of uniform mesopores. The exception is the catalyst 20Mo600, which contains almost only mesopores (the adsorption isotherm is only type IVa). All samples of prepared materials contain a mixture of micropores and mesopores, but the samples differ in the ratio of micropores and mesopores.
The total pore volume of pores less than 40 nm diameter at p/p° = 0.95 (Vtotal) of modified materials was found in the range from 0.184 cm3/g to 0.057 cm3/g, but the micropore volume (t-plot method) was in the range from 0.110 cm3/g to 0.002 cm3/g.
We observed sealing of pores connected with a decrease of specific surface area and decrease of micropore ratio with higher calcination temperature. Influence of temperature used during impregnation—room temperature or 60 °C did not influence textural characteristics of material (similar values in all measured parameters).
Nitrogen physisorption showed a significant decrease of catalyst specific surface area with using higher calcination temperature. This fact can also be visible from particle size distribution (Figure 5); in case of higher calcination temperatures, the increase of particle size is visible. Particles with larger size, present in 20Mo550, and 20Mo600 materials can be attributed to agglomerate formation using higher calcination temperatures. Non-modified BETA38 showed rather uniform distribution and can be classified as a monodisperse system, whereas modified materials are rather polydisperse systems.
In all cases, after modification of material with MoO3, particle sizes Dv(50) and Dv(90) increased compared to non-modified BETA38 material (Table 3). Materials prepared by wet impregnation at 60°C (20Mo450) and at room temperature (20Mo450RT) possess very similar Dv(50) and Dv(90)—1.64 and 1.94, and 35.2 and 44.2, respectively. A slight increase of value Dv(90) was observed when calcination temperature 500 °C was used (20Mo500). These results also corresponded to specific surface area results, where these materials had very similar characteristics. Another situation was observed using higher calcination temperatures—a significant increase in Dv(50) and Dv(90) values was observed—again, this is in accordance with physisorption measurement results, where a significant decrease of specific surface area was observed after calcining samples at 550 and 600 °C.
Morphology of materials surface was monitored using scanning electron microscopy (Figure 6). Elemental mapping confirmed the homogeneous distribution of molybdenum on the material surface, and no difference was observed between the materials. SEM images of materials BETA38, 20Mo450 and 20Mo600 are provided in supplementary material (Figures S1, S2 and S3).

2.2. Catalytic Testing

α-Pinene oxide isomerization can provide different products—among them, the most mentioned and valued are campholenic aldehyde and trans-carveol. Using our reaction system—molybdenum modified zeolite BETA38, the most abundant products were (Figure 7)—campholenic aldehyde (CA), trans-carveol (TCV), and p-methadien-2-ol (PMD). In some cases, p-cymene and sobreol were present in the reaction mixture (this diol occurs predominantly in the presence of water in a reaction system). The isomer of campholenic aldehyde—fencholenic aldehyde was in our reaction system present only in small amounts (< 5%), which was not surprising because its origination literature reported only using other types of catalyst (e.g., acid-treated clays [24]).
In the beginning, the activity of BETA38 (used as solid support of our catalysts) and activity of molybdenum-based catalysts prepared by different methods was compared (Figure 8).
The positive influence of molybdenum loading is visible—using unmodified zeolite BETA38 resulted in only low APO conversion (34%, 3 h) compared to Mo-modified materials (77−100%, 3 h). No significant difference was observed based on the preparation of materials either by wet impregnation at room temperature or by wet impregnation at 60 °C (both materials calcined at 450 °C). The selectivity of PMD formation differed only slightly (12% for RT and 14% for 60 °C), and this difference was only in the range of measurement error. On the other hand, the temperature of calcination after impregnation showed to be a crucial parameter—it significantly influenced APO conversion. Using the calcination temperatures 450 and 500 °C resulted in the same conversion of APO – 100% after 3 h. However, a further increase of calcination temperature led to APO conversion decrease (83% for 550 °C and 77% for 600 °C). A calcination temperature 550 °C caused a slight increase in CA selectivity. A significant loss of material surface area connected with higher calcination temperature was definitely the reason for the loss of material activity. We have chosen 20Mo450 material to test the influence of solvent type on reaction course (Table 4). Reaction course of reactions using different solvents are depicted in supplementary material (Figures S4–S6).
It is known that used solvent can significantly influence the reaction course; in our case of APO rearrangement, the selectivity to specific products would be the most important parameter. So far, to the best of our knowledge, widely studied solvent influence was performed using iron-based catalysts (Fe-zeolite, Fe-MCM-41) in APO rearrangement [15,21,41]. A wide number of solvents—nonpolar, polar protic, and polar aprotic—were used and further described using donor number to compare their Lewis basicity and dielectric constant to compare their polarity. Three nonpolar solvents were tested—toluene, cyclohexane, and 1,4-dioxane (Figure 9).
We obtained similar reaction results using nonpolar solvents comparing APO conversion and selectivity to single compounds (Figure 9). The most significant result can be observed in the initial reaction rate, which was using cyclohexane almost three times slower compared to toluene and 1,4-dioxane. On the other hand, the highest initial reaction rate was observed using 1,4-dioxane. In the case of nonpolar solvents, campholenic aldehyde was the major product, which is in accordance with the results already published in the literature [19,23,41]. As basicity cannot be the main reason (the difference between donor number of toluene and cyclohexane is negligible), other factors have to play an important role.
Usage of polar protic solvents—alcohols—was disadvantageous (Figure 10). Alcohols reacted with APO forming products with higher MW; it is likely that addition on a C-C double bond or oxirane ring-opening occurred. Nitromethane behaved differently probably (Table 4) because either by its different structure or by the fact that both its dielectric constant (35.9 vs. 15-21.8) and its donor number (2.7 vs. 19–25) are far different from dielectric constants and donor numbers of aforementioned alcohols. Using noncyclic alcohol, the total conversion of APO was obtained in the first 5 minutes of reaction.
We tested a large number of polar aprotic solvents following in this reaction (Table 4)—among them, the lowest APO conversions were obtained using pyridine and nitriles (acetonitrile, benzonitrile). In the case of pyridine, this fact was probably caused by its highest Lewis basicity from all studied solvents, which probably caused the blockage of acid sites present on catalyst surface. Pyridine is also known due to its strong adsorption on acid sites. It is the reason why this compound is used for basicity determination by different methods. Low catalytic activity in acetonitrile is in accordance with already published results using similar type of catalyst—Ti-BETA [16] and could be caused by its semi-basic character together with its high polarity. These both solvents provided low APO conversion (10−16%, 3 h), but high selectivity of CA formation (63−67%). Usage of N,N’-dimethylformamide offered only low APO conversion (27%, 3 h), which was surprising, as N,N’-dimethylacetamide provided almost twice higher APO conversion.
Polar aprotic solvents, which were able to provide APO conversion > 40% (3 h) and selectivity to sum of CA, TCV, and PMD > 50% (3 h) were chosen to evaluate the influence of dielectric constant and donor number on the reaction course.
The initial reaction rate (Figure 11) was independent on dielectric constant and on donor number, the highest values were obtained using ketones with mediate polarity and basicity from all studied solvents.
With the respect to APO conversion dependence on solvent polarity and basicity (Figure 12), the studied solvents were be divided into two groups, with except for nitromethane. Nitromethane did not fit in the studied influence of solvent donor number or dielectric constant on APO conversion. Its unique character compared to other studied aprotic solvents (low Lewis basicity and high polarity) could cause this. Excluding nitromethane, the first solvent group contained mediate polar and basic solvents (dielectric constant 5–20, donor number 3–20), and the second group – high polar and high basic solvents (dielectric constant 32–47, donor number 27–30 High APO conversion (96–100%, 3 h) was observed using the first solvent group. Among the second solvent group, significantly lower APO conversion was observed (41–45%, 3 h), which could probably be caused by the fact that basic solvents caused partial blockage of catalyst acid sites.
The solvent basicity significantly influenced the selectivity of CA and TCV (Figure 13), the effect of solvent polarity was not so straightforward. Overall, the selectivity of TCV formation increased with solvent basicity. On the other hand, the selectivity of CA formation decreased with increasing solvent basicity. Solvent basicity influenced CA and TCV formation considerably more than solvent polarity. The influence of solvent basicity on CA and TCV formation was in accordance with previously published results in the case of iron modified zeolite BETA or H3PW12O40 [14,23,41].
We observed similar CA selectivity (3 h) using both nitrobenzene and dichlorobenzene. It was probably caused by the fact, that these solvents possessed very similar Lewis basicity (donor numbers 2.7 and 3).
The selectivity of CA formation (Figure 14a) decreased with APO conversion in the case of dichlorobenzene, ethylacetate, N-methylpyrrolidone, and dimethylsulfoxide. In the case of other solvents, it remained the same in the range of measurement error. The selectivity of TCV formation (Figure 14b) was independent on APO conversion (was in the range of measurement error) in the case of most of the solvents except for N-methylpyrrolidone and dimethylsulfoxide. In the case of these two solvents, the selectivity of TCV formation increased significantly with increasing APO conversion. Especially DMSO was the most valuable solvent for the formation of TCV.
Based on literature, we proposed mechanism of two main products formation—CA and TCV (Figure 15).
Lewis acid site (LAS) activated epoxide ring in APO molecule inducing ring opening and formation of carbenium cation I. Cation I. might rearrange to secondary cyclopentanoic cation II. and tertiary p-menthenic cation III. Tertiary cation III. is more stable than secondary cation. Therefore, the formation of TCV should be thermodynamically controlled and CA formation should be kinetically controlled [19]. In the case of polar solvents, their solvation ability plays a role in cation I. stabilization, which would lead to the preferential formation of cation III. Moreover, in the case of the basic solvents, the proton transfer to cation III. can be assisted by the solvent, causing the preferential formation of TCV. CA formation from cation II. by C-C bond cleavage is then preferred in cases when the formation of cation III. is not favored, i.e., using nonpolar and nonbasic solvents [19].

3. Materials and Methods

Materials were prepared by wet impregnation from zeolite BETA (Si/Al = 38, CP814C, Zeolyst Int., Kansas city, KS, USA) and ammonium heptamolybdate tetrahydrate (Chemapol, Praha, Czech Republic, >99.5%). Two different wet impregnation processes were used: impregnation at room temperature [25,26,42] and at 60 °C [14]. To prepare the material with the desired Mo amount 20 wt.% 6.4 g of molybdate and 10 g of zeolite BETA was used. Different calcination temperatures were used (450−600°C). We denoted with the names containing calcination temperature and in the case of impregnation performed at room temperature also letters RT (Table 5).
In a typical reaction procedure, a round bottom flask was filled with solvent (6 ml), α-pinene oxide (1.25 g, Sigma Aldrich, Darmstadt, Germany, 97%), and the mixture was heated to 70 °C. Then catalyst (125 mg, calcined before reaction at 300 °C/air/4 h) was inserted into the reaction mixture, and the suspension was vigorously stirred for the next 3 hours at 70 °C. A wide number of solvents were used: 1,4-dioxane, N,N’-dimethylformamide, dimethyl sulfoxide, acetonitrile, cyclohexanone, cyclohexane, tetrahydrofuran, butan-1-ol, propan-1-ol, propan-2-ol, ethylacetate, toluene (all Penta, Praha, Czech Republic, >99.5%), butan-2-one, pyridine, butan-2-ol, (all Lachner, Neratovice, Czech Republic, > 99%), 1,2-dichlorobenzene, benzonitrile, cyclohexanol (all Sigma Aldrich, Darmstadt, Germany, 99%), nitromethane, N,N’-dimethylacetamide (both Sigma Aldrich, Darmstadt, Germany, 95%), N-methylpyrrolidone (Honeywell, Charlotte, NC, USA, 99%) and pentan-2-one (Aroma Praha, Židovice, Czech Republic, 95%). The reaction mixtures were analyzed using a gas chromatograph (Varian CP-3800, Palo Alto, CA, USA) equipped with a nonpolar column (DB-5) and FID, conversion (xAPO) and selectivity (s) were calculated according to the following equations. Representative chromatogram of reaction mixture is depicted at Figure S7.
x A P O = 100 r e l a t i v e   c o n c e n t r a t i o n A P O
s i = r e l a t i v e   c o n c e n t r a t i o n A P O x A P O
Prepared materials were characterized by X-Ray spectroscopy (X-Ray powder diffraction data were collected at room temperature with an X’Pert PRO θ–θ powder diffractometer with parafocusing Bragg–Brentano geometry using Cu Kα radiation (λ = 1.5418 Å, U = 40 kV, I = 30 mA) and X-ray fluorescence (an ARL 9400 XP sequential WD-XRF spectrometer was used to perform XRF analysis). Temperature programmed heating of non-calcined material was carried out using an AutoChem II Micromeritics Instrument 2920 (Unterschleissheim, Germany). Catalyst sample (0.1 g) was placed in a quartz U-shaped tube and during pretreatment the material was heated to 150 °C under argon flow (30 mL/min) and kept under this temperature for 1 h to remove physisorbed water. Afterward, the linear temperature program (10 °C min−1) started at a temperature of 150 °C and the sample was heated up in argon flow (30 mL/min) to a temperature of 700 °C. The specific surface area was determined using nitrogen adsorption (3Flex volumetric analyzer, Micromeritic, Unterschleissheim, Germany). The specific surface area of materials was calculated via the BET equation and t-plot method. The particle size distributions of all samples were determined by laser light scattering Malvern Mastersizer 3000 system equipped with Hydro EV wet sampling unit (Malvern Instruments Ltd., Cambridge, UK). The materials were characterized using a wet dispersion method. For each sample, particle size distributions were recorded for at least 5 determinations at an obscuration range of 8 to 14%. The particle size distributions of all materials were determined using demineralized water (conductivity < 1 µS·cm−1) with surfactant Triton X-100 (Aldrich, Darmstadt, Germany) as a dispersion medium. The reference refractive index for materials was 1.446. For SEM analysis, materials were applied on carbon adhesive tape and then goldened by 5 nm of gold via Quorum Q150 ES (Quorum Technologies Ltd., Laughton, UK). Scanning was performed by scanning electron microscope TESCAN VEGA 3 LMU (Tescan Brno, Czech Republic) in regime low vacuum (UniVac) at pressure in working chamber 1 Pa, accelerating voltage 20 kV and BSE mode. For elemental analysis EDS analyzer OXFORD Instrument X-max 20 mm2 with software Aztec (Oxford Instruments, Abingdon, UK) was used.

4. Conclusions

We demonstrated solvent influence on α-pinene oxide (APO) isomerization using molybdenum modified zeolite BETA. The type of used solvent had a significant impact on the reaction course. Nonpolar solvents provided > 95% APO conversion (3 h), and the major product was campholenic aldehyde. Usage of polar protic solvents was disadvantageous, as they reacted with the substrate preferentially. Among polar aprotic solvents, we observed more trends that are interesting. Basic solvents (dimethylsulfoxide, N,N’-dimethylacetamide, N,N’-dimethylformamide, N-methylpyrrolidone) provided significantly high selectivity of trans-carveol formation (>42%, 3 h) but at low APO conversion (27−45, 3 h). Using ketones, total APO conversion was obtained within 3 hours of the reaction, with the preferential formation of campholenic aldehyde (30−37% selectivity, 3 h). When using ethylacetate, dichlorobenzene, and nitromethane, high APO conversion (> 86%, 3 h), and preferential formation of campholenic aldehyde (37−45%) were observed. In nitriles, the highest selectivity of CA formation (~64%, 3 h) was obtained. However, APO conversion was very low (10−16%, 3 h). Pyridine showed to be an inappropriate solvent for this reaction. Only 3% APO conversion was present after 3 h. Its highest Lewis basicity from all studied solvents probably caused this fact. We showed that Mo modified zeolite beta is a possible available catalyst for the production of both campholenic aldehyde and trans-carveol. The modification of zeolite beta by molybdenum was advantageous for the reaction course. Moreover, we showed that the calcination temperature and solvent choice are critical parameters in the studied reaction.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/11/1244/s1, Figure S1: SEM images of BETA38, Figure S2: SEM images of 20Mo450, Figure S3: SEM images of 20Mo600, Figure S4: Reaction course using solvent toluene (catalyst 20Mo450, 70 °C), Figure S5: Reaction course using solvent nitromethane (catalyst 20Mo450, 70 °C), Figure S6: Reaction course using solvent dimethylsulfoxide (catalyst 20Mo450, 70 °C), Figure S7: Representative chromatogram of reaction mixture (solvent toluene).

Author Contributions

Conceptualization, E.V. (Eva Vrbková) and E.V. (Eliška Vyskočilová); methodology, E.V. (Eva Vrbková) and M.L.; formal analysis, E.V. (Eva Vrbková) and M.L.; writing—original draft preparation, E.V. (Eva Vrbková); writing—review and editing, E.V. (Eliška Vyskočilová); supervision, L.Č.; project administration, E.V. (Eva Vrbková). All authors have read and agreed to the published version of the manuscript.

Funding

This work was realized within the Operational Programme Prague−Competitiveness (CZ.2.16/3.1.00/24501) and “National Program of Sustainability” ((NPU I LO1613) MSMT-43760/2015) and from Institutional funding of UCT Prague. The APC was funded from Institutional funding of UCT Prague.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Edl, W.; Sienel, G.R. Process for the production of epoxides. U.S. Patent No. 4882442A, 30 December 1989. [Google Scholar]
  2. Jaafari, A.; Tilaoui, M.; Mouse, H.A.; M’bark, L.A.; Aboufatima, R.; Chait, A.; Lepoivre, M.; Zyad, A. Comparative study of the antitumor effect of natural monoterpenes: Relationship to cell cycle analysis. Revista Brasileira Farmacognosia 2012, 22, 534–540. [Google Scholar] [CrossRef] [Green Version]
  3. Hussain, A.I.; Anwar, F.; Shahid, M.; Ashraf, M.; Przybylski, R. Chemical Composition, and Antioxidant and Antimicrobial Activities of Essential Oil of Spearmint (Mentha spicata L.). J. Essent. Oil Res. 2010, 22, 78–84. [Google Scholar] [CrossRef]
  4. Castro, J.M.; Linares-Palomino, P.J.; Salido, S.; Altarejos, J.; Nogueras, M.; Sánchez, A. Enantiospecific synthesis, separation and olfactory evaluation of all diastereomers of a homologue of the sandalwood odorant Polysantol. Tetrahedron 2005, 61, 11192–11203. [Google Scholar] [CrossRef]
  5. Arbushow, B. Studium der Isomerisation von Terpen-oxyden, I. Mitteil.: Isomerisation des α-Pinen-oxydes bei der Reaktion von Reformatsky. Chem. Ber. 1935, 68, 1430–1435. [Google Scholar] [CrossRef]
  6. Duetz, W.; Fjällman, A.; Ren, S.; Jourdat, C.; Witholt, B. Biotransformation of D-Limonene to (+) trans-Carveol by Toluene-Grown Rhodococcus opacus PWD4 Cells. Appl. Environ. Microbiol. 2001, 67, 2829–2832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Coelho, J.V.; de Meireles, A.L.P.; da Silva Rocha, K.A.; Pereira, M.C.; Oliveira, L.C.A.; Gusevskaya, E.V. Isomerization of α-pinene oxide catalyzed by iron-modified mesoporous silicates. Appl. Catal. A Gen. 2012, 443–444, 125–132. [Google Scholar] [CrossRef]
  8. Kumar, N.; Mäki-Arvela, P.; Diaz, S.F.; Aho, A.; Demidova, Y.; Linden, J.; Shepidchenko, A.; Tenhu, M.; Salonen, J.; Laukkanen, P.; et al. Isomerization of α-Pinene Oxide Over Iron-Modified Zeolites. Top. Catal. 2013, 56, 696–713. [Google Scholar] [CrossRef]
  9. Ravasio, N.; Zaccheria, F.; Gervasini, A.; Messi, C. A new, Fe based, heterogeneous Lewis acid: Selective isomerization of α-pinene oxide. Catal. Commun. 2008, 9, 1125–1127. [Google Scholar] [CrossRef]
  10. Štekrová, M.; Kumar, N.; Aho, A.; Sinev, I.; Grunert, W.; Dahl, J.; Roine, J.; Arzumanov, S.S.; Mäki-Arvela, P.; Murzin, D.Y. Isomerization of alpha-pinene oxide using Fe-supported catalysts: Selective synthesis of campholenic aldehyde. Appl. Catal. A Gen. 2014, 470, 162–176. [Google Scholar] [CrossRef]
  11. Sánchez-Velandia, J.E.; Villa, A.L. Isomerization of α- and β-pinene epoxides over Fe or Cu supported MCM-41 and SBA-15 materials. Appl. Catal. A Gen. 2019, 580, 17–27. [Google Scholar] [CrossRef]
  12. Dhakshinamoorthy, A.; Alvaro, M.; Chevreau, H.; Horcajada, P.; Devic, T.; Serre, C.; Garcia, H. Iron(iii) metal–organic frameworks as solid Lewis acids for the isomerization of α-pinene oxide. Catal. Sci. Technol. 2012, 2, 324–330. [Google Scholar] [CrossRef]
  13. Fellenz, N.A.; Bengoa, J.F.; Marchetti, S.G.; Gervasini, A. Influence of the Brönsted and Lewis acid sites on the catalytic activity and selectivity of Fe/MCM-41 system. Appl. Catal. A Gen. 2012, 435-436, 187–196. [Google Scholar] [CrossRef]
  14. Vyskočilová, E.; Hašková, L.; Červený, L. Solvent-induced selectivity in α-pinene oxide isomerization catalyzed by Fe-modified zeolite beta. Chem. Pap. 2019, 73, 1621–1627. [Google Scholar] [CrossRef]
  15. Štekrová, M.; Kumar, N.; Diaz, S.F.; Mäki-Arvela, P.; Murzin, D.Y. H- and Fe-modified zeolite beta catalysts for preparation of trans-carveol from α-pinene oxide. Catal. Today 2015, 241, 237–245. [Google Scholar] [CrossRef]
  16. Kunkeler, P.J.; van der Waal, J.C.; Bremmer, J.; Zuurdeeg, B.J.; Downing, R.S.; van Bekkum, H. Application of zeolite titanium Beta in the rearrangement of α-pinene oxide to campholenic aldehyde. Catal. Lett. 1998, 53, 135–138. [Google Scholar] [CrossRef]
  17. Panadero, M.P.; Velty, A. Readily available Ti-beta as an efficient catalyst for greener and sustainable production of campholenic aldehyde. Catal. Sci. Technol. 2019, 9, 4293–4303. [Google Scholar] [CrossRef]
  18. Pitínová-Štekrová, M.; Eliášová, P.; Weissenberger, T.; Shamzy, M.; Musilová, Z.; Čejka, J. Highly selective synthesis of campholenic aldehyde over Ti-MWW catalysts by α-pinene oxide isomerization. Catal. Sci. Technol. 2018, 8, 4690–4701. [Google Scholar] [CrossRef]
  19. Costa, V.V.; da Silva Rocha, K.A.; de Sousa, L.F.; Robles-Dutenhefner, P.A.; Gusevskaya, E.V. Isomerization of α-pinene oxide over cerium and tin catalysts: Selective synthesis of trans-carveol and transsobrerol. J. Mol. Catal. A Chem. 2011, 345, 69–74. [Google Scholar] [CrossRef] [Green Version]
  20. Štekrová, M.; Matoušková, M.; Vyskočilová, E.; Červený, L. Selective preparation of campholenic aldehyde over heterogenized methyltrioxorhenium. Res. Chem. Intermed. 2015, 41, 9003–9013. [Google Scholar] [CrossRef]
  21. Štekrová, M.; Kumar, N.; Mäki-Arvela, P.; Ardashov, O.V.; Volcho, K.P.; Salakhutdinov, N.F.; Murzin, D.Y. Selective Preparation of trans-Carveol over Ceria Supported Mesoporous Materials MCM-41 and SBA-15. Materials 2013, 6, 2103–2118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Da Silva Rocha, K.A.; Kozhevnikov, I.V.; Gusevskaya, E.V. Isomerisation of α-pinene oxide over silica supported heteropoly acid H3PW12O40. Appl. Catal. A 2005, 294, 106–110. [Google Scholar] [CrossRef]
  23. Da Silva Rocha, K.A.; Hoehne, J.L.; Gusevskaya, E.V. Phosphotungstic Acid as a Versatile Catalyst for the Synthesis of Fragrance Compounds by α-Pinene Oxide Isomerization: Solvent-Induced Chemoselectivity. Chem. Eur. J. 2008, 14, 6166–6172. [Google Scholar] [CrossRef] [PubMed]
  24. Sidorenko, A.Y.; Ignatovich, Z.V.; Ermolinskaya, A.L.; Kravtsova, A.V.; Baranovskii, A.V.; Koroleva, E.V.; Agabekov, V.E. Synthesis of Fencholenic Aldehyde from α-pinene Epoxide on Modified Clays. Chem. Nat. Compd. 2018, 54, 893–897. [Google Scholar] [CrossRef]
  25. Vrbková, E.; Šteflová, B.; Sekerová, L.; Krupka, J.; Vyskočilová, E.; Červený, L. Contribution to MoO 3–SiO 2 and WO 3–SiO 2 utilization—Active catalysts in jasmine aldehyde, 2-hexyl-1, 3-dioxolane and methyllaurate synthesis. React. Kinet. Mech. Catal. 2020, 129, 645–658. [Google Scholar] [CrossRef]
  26. Vrbková, E.; Šteflová, B.; Vyskočilová, E.; Červený, L. Heterogeneous Mo/W/Zn–SiO2 based catalysts in nopol (2-(6,6-dimethyl-2-bicyclo[3.1.1]hept-2-enyl)ethanol) synthesis. React. Kinet. Mech. Catal. 2020, 131, 213–232. [Google Scholar] [CrossRef]
  27. Wehrer, W.; Bigey, L.; Hilaire, L. Catalytic reactions of n-hexane and 1-hexene on molybdenum dioxide. Appl. Catal. A 2003, 243, 109–119. [Google Scholar] [CrossRef]
  28. Bouchy, C.; Pham-Huu, C.; Heinrich, B.; Chaumont, C.; Ledoux, M.J. Microstructure and Characterization of a Highly Selective Catalyst for the Isomerization of Alkanes: A Molybdenum Oxycarbide. J. Catal. 2000, 190, 92–103. [Google Scholar] [CrossRef]
  29. Lamic, A.F.; Shin, C.H.; Djega-Mariadassou, G.; Potvin, C. Characterization of Mo2C–WO2 composite catalysts for bifunctional isomerization: A new pulse method to quantify acid sites. Appl. Catal. A 2006, 302, 5–13. [Google Scholar] [CrossRef]
  30. York, A.P.E.; Pham-Huu, C.; Del Gallo, P.; Ledoux, M.J. Molybdenum oxycarbide hydrocarbon isomerization catalysts: Cleaner fuels for the future. Catal. Today 1997, 35, 51–57. [Google Scholar] [CrossRef]
  31. Ohno, T.; Li, Z.; Sakai, N.; Sakagami, H.; Takahashi, N.; Matsuda, T. Heptane isomerization over molybdenum oxides obtained by H2 reduction of HxMoO3 with different hydrogen contents. Appl. Catal. A 2010, 389, 52–59. [Google Scholar] [CrossRef]
  32. Bruno, S.M.; Valente, A.A.; Pillinger, M.; Amelse, J.; Romão, C.C.; Gonçalves, I.S. Efficient Isomerization of α-Pinene Oxide to Campholenic Aldehyde Promoted by a Mixed-Ring Analogue of Molybdenocene. ACS Sustain. Chem. Eng. 2019, 7, 13639–13645. [Google Scholar] [CrossRef]
  33. Caullet, P.; Hazm, J.; Guth, J.; Joly, J.; Lynch, J.; Raatz, F. Synthesis of zeolite Beta from nonalkaline fluoride aqueous aluminosilicate gels. Zeolites 1992, 12, 240–250. [Google Scholar] [CrossRef]
  34. Hansen, S.; Andersson, A. Electron microscopy of some molybdenum oxide phases after use as catalysts in oxidative ammonolysis and ammoxidation of toluene. J. Solid State Chem. 1988, 75, 225–243. [Google Scholar] [CrossRef]
  35. Sekerová, L.; Vyskočilová, E.; Červený, L. Prins cyclization of isoprenol with various aldehydes using MoO3/SiO2 as a catalyst. React. Kinet. Mech. Catal. 2017, 121, 83–95. [Google Scholar] [CrossRef]
  36. Gutmann Acceptor and Donor number. Available online: http://www.stenutz.eu/chem/solv21.php (accessed on 12 August 2020).
  37. Solvent Properties Chart. Available online: https://depts.washington.edu/eooptic/linkfiles/dielectric_chart%5B1%5D.pdf (accessed on 12 August 2020).
  38. Munack, A.; Schmidt, L.; Schröder, O.; Schaper, K.; Pabst, C.; Krahl, J. Alcohols as a means to inhibit the formation of precipitates in blends of biodiesel and fossil diesel fuel. Agric. Eng. Int. 2015, 2015, 226–233. [Google Scholar]
  39. Dielectric Constants of Liquids. Available online: https://www.engineeringtoolbox.com/liquid-dielectric-constants-d_1263.html (accessed on 31 August 2020).
  40. Dielectric Constants of Common Materials. Available online: https://www.kabusa.com/Dilectric-Constants.pdf (accessed on 31 August 2020).
  41. Mäki-Arvela, P.; Shcherban, N.; Lozachmeur, C.; Russo, V.; Wärnå, J.; Murzin, D.Y. Isomerization of α-Pinene Oxide: Solvent Effects, Kinetics and Thermodynamics. Catal. Lett. 2019, 149, 203–214. [Google Scholar] [CrossRef] [Green Version]
  42. Vrbková, E.; Šteflová, B.; Zapletal, M.; Vyskočilová, E.; Červený, L. Tungsten oxide-based materials as effective catalysts in isopulegol formation by intramolecular Prins reaction of citronellal. Res. Chem. Intermed. 2020, 46, 4047–4059. [Google Scholar] [CrossRef]
Figure 1. Structure of trans-carveol (a), campholenic aldehyde (b) and sandalore (c).
Figure 1. Structure of trans-carveol (a), campholenic aldehyde (b) and sandalore (c).
Catalysts 10 01244 g001
Figure 2. TCD (thermal conductivity detector) signal from temperature programmed heating of noncalcined sample.
Figure 2. TCD (thermal conductivity detector) signal from temperature programmed heating of noncalcined sample.
Catalysts 10 01244 g002
Figure 3. X-Ray diffractograms of nonmodified BETA38 and molybdenum modified materials (° alumisilicate bands, * Mo bands).
Figure 3. X-Ray diffractograms of nonmodified BETA38 and molybdenum modified materials (° alumisilicate bands, * Mo bands).
Catalysts 10 01244 g003
Figure 4. Adsorption isotherms of nonmodified BETA38 and molybdenum modified materials.
Figure 4. Adsorption isotherms of nonmodified BETA38 and molybdenum modified materials.
Catalysts 10 01244 g004
Figure 5. Particle size distribution of nonmodified BETA38 and molybdenum modified materials.
Figure 5. Particle size distribution of nonmodified BETA38 and molybdenum modified materials.
Catalysts 10 01244 g005
Figure 6. SEM image of 20Mo450 (a) and 20Mo600 (b) material with elemental maps of Si and Mo.
Figure 6. SEM image of 20Mo450 (a) and 20Mo600 (b) material with elemental maps of Si and Mo.
Catalysts 10 01244 g006
Figure 7. Reaction scheme: campholenic aldehyde (a), fencholenic aldehyde (b), trans-carveol (c), p-methadien-2-ol (d), p-cymene (e), and sobreol (f).
Figure 7. Reaction scheme: campholenic aldehyde (a), fencholenic aldehyde (b), trans-carveol (c), p-methadien-2-ol (d), p-cymene (e), and sobreol (f).
Catalysts 10 01244 g007
Figure 8. Reaction result using zeolite BETA38 and materials modified with 20 wt.% of molybdenum prepared by different ways (1.25 g APO, 6 mL of toluene, 0.125 g of catalyst, 70 °C, 3 h).
Figure 8. Reaction result using zeolite BETA38 and materials modified with 20 wt.% of molybdenum prepared by different ways (1.25 g APO, 6 mL of toluene, 0.125 g of catalyst, 70 °C, 3 h).
Catalysts 10 01244 g008
Figure 9. Reaction result using catalyst 20Mo450 and different nonpolar solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst, 70 °C, 3 h).
Figure 9. Reaction result using catalyst 20Mo450 and different nonpolar solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst, 70 °C, 3 h).
Catalysts 10 01244 g009
Figure 10. Reaction result using catalyst 20Mo450 and different polar protic solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst, 70 °C, 3 h).
Figure 10. Reaction result using catalyst 20Mo450 and different polar protic solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst, 70 °C, 3 h).
Catalysts 10 01244 g010
Figure 11. Initial reaction rate using different polar aprotic solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst 20Mo450, 70 °C, 3 h).
Figure 11. Initial reaction rate using different polar aprotic solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst 20Mo450, 70 °C, 3 h).
Catalysts 10 01244 g011
Figure 12. APO conversion using different polar aprotic solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst 20Mo450, 70 °C, 3 h), dichlorobenzene (DCB), butan-2-one (BN), cyclohexanone (CYN), N,N’-dimethylacetamide (DMA), dimethylsulfoxide (DMSO), ethylacetate (EtOAc), N-methylpyrrolidone (NMP), nitromethane (NM), pentan-2-one (PN), tetrahydrofuran (THF).
Figure 12. APO conversion using different polar aprotic solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst 20Mo450, 70 °C, 3 h), dichlorobenzene (DCB), butan-2-one (BN), cyclohexanone (CYN), N,N’-dimethylacetamide (DMA), dimethylsulfoxide (DMSO), ethylacetate (EtOAc), N-methylpyrrolidone (NMP), nitromethane (NM), pentan-2-one (PN), tetrahydrofuran (THF).
Catalysts 10 01244 g012
Figure 13. Selectivity to CA (black square), TCV (red circle) and PMD (blue triangle) using different polar aprotic solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst 20Mo450, 70 °C, 3 h).
Figure 13. Selectivity to CA (black square), TCV (red circle) and PMD (blue triangle) using different polar aprotic solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst 20Mo450, 70 °C, 3 h).
Catalysts 10 01244 g013
Figure 14. Dependence on campholenic aldehyde (CA, a) and trans-carveol (TCV, b) formation selectivity on APO conversion using different polar aprotic solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst 20Mo450, 70 °C) dichlorobenzene (DCB), butan-2-one (BN), cyclohexanone (CYN), N,N’-dimethylacetamide (DMA), dimethylsulfoxide (DMSO), ethylacetate (EtOAc), N-methylpyrrolidone (NMP), nitromethane (NM), pentan-2-one (PN), tetrahydrofuran (THF).
Figure 14. Dependence on campholenic aldehyde (CA, a) and trans-carveol (TCV, b) formation selectivity on APO conversion using different polar aprotic solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst 20Mo450, 70 °C) dichlorobenzene (DCB), butan-2-one (BN), cyclohexanone (CYN), N,N’-dimethylacetamide (DMA), dimethylsulfoxide (DMSO), ethylacetate (EtOAc), N-methylpyrrolidone (NMP), nitromethane (NM), pentan-2-one (PN), tetrahydrofuran (THF).
Catalysts 10 01244 g014
Figure 15. Proposed mechanism for CA and TCV formation; LAS represents Lewis acid sites of MoO3.
Figure 15. Proposed mechanism for CA and TCV formation; LAS represents Lewis acid sites of MoO3.
Catalysts 10 01244 g015
Table 1. X-ray fluorescence (XRF) analysis results.
Table 1. X-ray fluorescence (XRF) analysis results.
MaterialMoO3 (%)Mo (%) calculatedSiO2 (%)Al2O3 (%)Others (%)
20Mo45032.821.663.73.40.1
20Mo450RT36.023.760.83.10.1
20Mo50032.021.164.73.20.1
20Mo55033.221.963.53.20.1
20Mo60033.922.362.93.10.1
Table 2. Textural characteristic of prepared materials and unmodified BETA38 zeolite.
Table 2. Textural characteristic of prepared materials and unmodified BETA38 zeolite.
MaterialSBET m2/gSt-plot m2/gTotal Pore Volume cm3/gt-Plot Micropore Volume cm3/gRatio of Micropores
(%)
BETA38539.62136.150.3410.21061.6
20Mo450272.8158.700.1810.11060.8
20Mo500239.4653.980.1710.09656.1
20Mo550154.3343.470.1390.05741.0
20Mo60016.5012.430.0570.0023.5
20Mo450RT269.3061.610.1840.10758.1
Table 3. Particle characteristic of prepared materials.
Table 3. Particle characteristic of prepared materials.
MaterialSpan ValueParticle Size (µm)
Dv(10)Dv(50)Dv(90)
BETA381.60.390.711.6
20Mo45022.20.431.6435.2
20Mo450RT22.50.441.9444.2
20Mo50047.90.461.7455.0
20Mo55081.90.504.74389.0
20Mo60043.80.5816.20710.0
Table 4. Reaction results using different solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst 20Mo450, 70 °C, 3 h) [36,37,38,39,40].
Table 4. Reaction results using different solvents (1.25 g APO, 6 mL of solvent, 0.125 g of catalyst 20Mo450, 70 °C, 3 h) [36,37,38,39,40].
SolventSolvent TypeDonor Number (kJ/mol)Dielectric Constant (Relative Permitivity)Initial Reaction Rate (mmol/gcat.min)APO Conversion at 3 h (%)CA Selectivity at 3 h (%)TCA Selectivity at 3 h (%)PMD Selectivity at 3 h (%)
cyclohexanenonpolar02.022.409934.617.413.9
toluenenonpolar0.12.385.4310034.214.814.0
1,4 dioxanenonpolar14.32.256.6310028.718.011.0
nitromethanepolar aprotic2.735.874.688644.910.712.3
butan-1-olpolar protic19.517.813.111003.27.01.9
propan-1-olpolar protic19.821.813.111003.36.42.7
propan-2-olpolar protic21.117.913.1110018.214.25.4
dichlorbenzenepolar aprotic39.932.6710042.113.313.1
benzonitrilepolar aprotic11.9260.711666.59.80.0
acetonitrilepolar aprotic1437.50.681062.74.92.9
ethylacetatepolar aprotic17.16.022.679737.115.813.9
pentan-2-onepolar aprotic17.118.211.3210035.115.512.0
butan-2-onepolar aprotic17.418.511.9510030.810.49.1
cyclohexanonepolar aprotic1818.28.3410037.415.016.6
tetrahydrofuranpolar aprotic207.586.7410031.728.96.5
cyclohexanolpolar protic25155.5810024.121.15.1
N,N’-dimethylformamidepolar aprotic26.636.71.012726.543.814.0
N-methylpyrrolidonepolar aprotic27.332.170.854126.142.912.2
N,N’-dimethylacetamidepolar aprotic27.837.84.654425.345.511.5
dimethylsulfoxidepolar aprotic29.846.70.974525.353.612.7
pyridinepolar aprotic33.112.40.14355.618.50.0
Table 5. Denotation of prepared materials.
Table 5. Denotation of prepared materials.
Material DenotationTemperature of Wet ImpregnationCalcination Temperature (°C)
20Mo450RTroom temperature450
20Mo45060 °C450
20Mo500500
520Mo550550
20Mo600600
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Vrbková, E.; Vyskočilová, E.; Lhotka, M.; Červený, L. Solvent Influence on Selectivity in α-Pinene Oxide Isomerization Using MoO3-Modified Zeolite BETA. Catalysts 2020, 10, 1244. https://doi.org/10.3390/catal10111244

AMA Style

Vrbková E, Vyskočilová E, Lhotka M, Červený L. Solvent Influence on Selectivity in α-Pinene Oxide Isomerization Using MoO3-Modified Zeolite BETA. Catalysts. 2020; 10(11):1244. https://doi.org/10.3390/catal10111244

Chicago/Turabian Style

Vrbková, Eva, Eliška Vyskočilová, Miloslav Lhotka, and Libor Červený. 2020. "Solvent Influence on Selectivity in α-Pinene Oxide Isomerization Using MoO3-Modified Zeolite BETA" Catalysts 10, no. 11: 1244. https://doi.org/10.3390/catal10111244

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