Next Article in Journal
Applicability of Nickel-Based Catalytic Systems for Hydrodehalogenation of Recalcitrant Halogenated Aromatic Compounds
Previous Article in Journal
Modification of Silica Xerogels with Polydopamine for Lipase B from Candida antarctica Immobilization
Previous Article in Special Issue
Transition Metal Ions as Ozonation Catalysts: An Alternative Process of Heterogeneous Catalytic Ozonation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Brominated Alkanes via Heterogeneous Catalytic Distillation over Al2O3/SO42−/ZrO2

School of Chemistry and Chemical Engineering, Southeast University, Nanjing 211189, China
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(12), 1464; https://doi.org/10.3390/catal11121464
Submission received: 4 November 2021 / Revised: 28 November 2021 / Accepted: 29 November 2021 / Published: 30 November 2021
(This article belongs to the Special Issue Selectivity and Stability of Heterogeneous Catalysts)

Abstract

:
Concentrated sulfuric acid is generally used as a catalyst for producing brominated alkanes in traditional methods, but is highly corrosive and difficult to separate. This work reports the preparation of bromopropane from n-propanol based on a reactive distillation strategy combined with alumina-modified sulfated zirconia (Al2O3/SO42−/ZrO2) as a heterogenous catalyst. As expected, under the optimum reaction conditions (110 °C), the yield of bromopropane was 96.18% without side reactions due to the reactive distillation strategy. Meanwhile, the microscopic morphology and performance of Al2O3/SO42−/ZrO2 were evaluated by X-ray diffraction (XRD), scanning electron microscopy (SEM), X-ray photoelectron spectroscopy (XPS), Brunner–Emmet–Teller (BET), Fourier transform infrared spectroscopy (FT–IR), and other characterization methods. The results confirmed that the morphology of zirconia sulfate is effectively regulated by the modification method of alumina, and more edges and angles provide more catalytic acid sites for the reaction. Furthermore, Al2O3/SO42−/ZrO2 exhibited high stability and remarkable reusability due to the strong chemical bond Zr–Al–Zr. This work provides a practical method for the preparation of bromopropane and can be further extended to the preparation of other bromoalkanes.

Graphical Abstract

1. Introduction

The development of green synthesis technology and improvement of atomic economic efficiency have become hotspots of current research with the trend toward increasing industrialization and urbanization. Brominated alkanes are critical industrial solvents and raw chemical materials, used as intermediates for various products such as medicines, pesticides, perfumes, dyes, surfactants, and photosensitive materials. However, the traditional method uses sulfuric acid as a catalyst in brominated alkane production, resulting in several tough issues including equipment corrosion, by-product separation, and environmental pollution. Taking the factors mentioned above into account, developing a low-cost, pragmatic, and efficient method for producing brominated alkanes remains a challenge.
As early as the middle of the 19th century, scientists began to research solid acid catalysts. Currently, several solid acid catalysts are used in various catalytic reactions and industry. Compared to liquid acid catalysts, solid catalysts have multiple advantages including a simple preparation process, high reactivity, and low corrosivity. Furthermore, the use of solid acid catalysts also alleviates various problems caused by homogeneous reactions, such as the separation and recovery of products and catalysts. Sulfated zirconium oxide (SZ) is one of the metal oxide solid acid catalysts widely studied this year due to its high reactivity and environmental friendliness. It has a wide range of applications in alkylation, isomerization, and other reactions [1,2,3]. In 1979, Hino et al. [4]. synthesized a sulfated oxide super acid catalyst, SO42−/ZrO2, and several studies since have thoroughly researched its preparation process and other aspects. However, SZ does have some issues including fast deactivation, poor channel structure, and insufficient acid sites. In early research, scientists modified sulfated zirconium oxide with expensive metals such as platinum and palladium to improve its catalytic activity [5].
As one of the most used additives and catalytic aids, alumina can improve the catalyst structure and properties in several ways, including mechanical strength and catalyst formation, catalyst-specific surface properties, acidic properties, reactivity, etc. [6]. At the same time, it is also low-cost, easily accessed, conveniently stored, and environment-friendly. In the field of sulfurized zirconia, several studies have indicated that adding a small amount of alumina component to the catalyst sample can be advantageous in producing a crystalline structure, aperture structure, surface sulfate load, catalyst acidity, and structural stability [7]. In a light alkane isomerization reaction, alumina added to super acid typically produces a significantly greater reaction conversion rate and higher stability [8,9]. A large body of literature has reported on the modification effect of Al2O3 on superacid catalysts. Kim et al. [10] found that alumina-modified super acid increased the reaction conversion rate by about 30% in butane isomerization reactions. At the same time, the addition of aluminum would also change the composition and properties of the original acidic catalyst center. Zhou et al. [3] used a variety of laboratory and industrial catalyst preparation methods to prepare alumina-modified solid composite catalysts. The results showed that the addition of active aluminum significantly impacts the structural and acidic properties of the catalyst. Cao [11,12] used aluminum sulfate and gallium sulfate as vulcanizing agents when introducing sulfur-containing species into the catalyst preparation process. Compared with the traditional SO42−/ZrO2 catalyst, this approach improved the structural performance of the catalytic material while increasing activity, and delayed the growth of ZrO2 crystal grains.
Under normal circumstances, metal oxides and SO42− have three coordination forms: single coordination, chelating double coordination, and bridge double coordination [13]. In the SO42−/MxOy type solid super acid, the latter two coordination effects between SO42− and metal oxide dominate [14]. The acidic center of the catalyst super acid is due to the strong induction effect between the metal oxide and SO42−. During the calcination process, the ionic S=O bond is transformed into a covalent S=O bond, and the S=O group is a strong electron withdrawing group. Due to the electron-withdrawing induction effect of the covalent bond, the electron cloud density on the M–O bond is reduced, resulting in a strong L acid center, which then exhibits super acidity, and water molecules or –OH adsorb and dissociate at the L acid center. The B acid center is also produced. The synergistic effect of B and L acid centers makes the acidity of the catalyst higher than that of 100% sulfuric acid. The reaction mechanism is shown in Figure S1.
Inspired by the above works, alumina modified sulfated zirconia (Al2O3/SO42−/ZrO2) through co-precipitation was used as a heterogeneous catalyst to replace the traditional liquid acid catalyst for bromination reaction. The optimal preparation conditions for Al2O3/SO42−/ZrO2 were obtained by screening the catalyst preparation and reaction conditions. Meanwhile, the reactive distillation strategy was proposed for the first time in this reaction, in which both separation and rectification are carried out simultaneously for rapid product transfer. Rapid removal of the product not only overcomes the limit of equilibrium conversion of reversible reaction, but also reduces energy consumption. Therefore, the research content of this article has important significance and reference value for a green brominated alkane production process and the application of Al2O3 modified SO42−/ZrO2 solid acid catalysts. Furthermore, a catalytic distillation reaction was used to separate and prepare bromopropane to reduce energy consumption during product purification, contributing to the development prospects of current green chemistry.

2. Results

2.1. Characterization of the Catalysts

Figure 1A displays the XRD spectra of Al2O3/SO42−/ZrO2 catalysts at different calcination temperatures. It can be seen from Figure 1A that when the calcination temperature was lower than 600 °C, there was no obvious characteristic peak. The reason for this was deemed to be that the calcination temperature was too low, and no crystals were formed in the catalyst, thus 600 °C was selected as the calcination temperature of the catalyst.
Figure 1B displays the XRD spectra of Al2O3/SO42−/ZrO2 catalysts with different Al2O3 loadings, at a calcination temperature of 600 °C. It can be seen from Figure 1B that the diffraction peaks 2θ = 30.5, 35.4, 50.9, and 60.8° were found in almost all samples. After comparison with relevant literature [10,15], the results were verified to be the diffraction peaks of ZrO2 crystal, indicating that the catalyst support of SO42−/ZrO2 had been successfully prepared. The literature detailed a catalyst modified by Al2O3, and which predominantly consisted of tetragonal phase type ZrO2 and a small amount of monoclinic phase type ZrO2 [16]. The sample XRD spectra of a small amount of Al2O3 added via co-precipitation shown that it completely changed to tetragonal phase, as the monoclinic phase type ZrO2 disappeared. Zalewski et al. [17] believe that in sulfurized zirconia catalysts, the modest amount of small Al2O3 grains added via co-precipitation was mainly concentrated in the ZrO2lattice gaps, causing the ZrO2 crystal grain size to be less than the pure sulfurized zirconia after calcination. At the same time, the small ZrO2 crystal grains made it easier to form a metastable tetragonal phase type ZrO2 during the calcination process [18]. Relevant literature shows that additional tetragonal crystal ZrO2 is beneficial in ensuring the excellent catalytic activity of sulfated zirconia catalysts [19]. There is a corresponding relationship between catalytic activity and the proportion of tetragonal crystal [20]. There was no Al2O3 diffraction peak in the spectrum, which means that Al2O3 was uniformly dispersed in the sample and the crystal size was very small.
A scanning electron microscope was used to further observe and analyze the ZrO2 catalyst morphology pre- and post- modification with Al2O3, as shown in Figure 2A,B. The results showed that the sulfated zirconia catalyst particles were not modified by aluminum and were large and irregularly arranged with a relatively rough surface, which provides more acidic sites for catalytic reactions. The surface morphology of the composite catalyst modified by co-precipitation had undergone significant changes. It can be seen from Figure 2B that the surface of the catalyst particles became smoother, the particles became smaller, and their surfaces became denser. The surface morphology of the catalyst changed because ZrO2 and alumina are relatively tightly bonded during the preparation process, and are likely to affect each other during the high-temperature calcination process. Furthermore, alumina, as a relatively rigid oxide, easily forms a specific structure and surface morphology at high temperature, which will impact the specific surface properties of the catalyst.
BET technology was used to study the effect of alumina addition on the catalyst specific surface area, and the results are shown in Table 1. It can be seen that with the addition of alumina, the catalyst overall specific surface area showed an increasing trend, and pore diameter decreased. This is because a microporous alumina system was introduced into a mesoporous ZrO2 system. It is generally believed that the larger the catalyst-specific surface area, the more active sites are exposed, which enhances the reaction progress.
Additionally, the Al2O3 modified ZrO2 catalyst was subjected to energy–dispersive X-ray spectroscopy (EDS) element distribution measurement, and a catalyst sample with a Zr/Al ratio of 4:1 was selected for measurement and analysis, as shown in Figure 3. EDS element analysis showed that the element composition was slightly different from the original, by about 4.1. The deviation analysis considered element loss during synthesis, and a spectrum analysis was obtained after three repetitions. It can be seen from the figure that Al was uniformly distributed in the material, and the S content was 3.37%.

2.2. Catalytic Activity of Al2O3/SO42−/ZrO2

To determine the influence of alumina content on the binding state of sulfate radicals on the catalyst, FT-IR characterization was performed on catalysts with different alumina contents. The structure is shown in Figure 4. According to the relevant literature [21,22], the bands at 1120 and 1240 cm−1 were attributable to the asymmetric O=S=O stretching vibration [23,24], and the band at 1640 cm−1 was attributable to the deformation vibration of adsorbed water. The figure shows that as the alumina content increased, the O=S=O stretching vibration band gradually shifted, and the state of the surface catalyst sulfate species was affected by the catalyst content.
To further determine the valence state of the elements in the Al2O3/SO42−/ZrO2 catalyst, the catalyst was characterized by XPS, as shown in Figure 5. The spectrum indicated the presence of Zr, Al, C, and S. According to Figure 5 and related literature, the S 2p region of the Al2O3/SO42−/ZrO2 catalyst had only one peak at 169 eV [25,26], which belongs to a positive hexavalent sulfate species, indicating that only the highest valence sulfate was present in the catalyst.
We now know that the Al2O3 component in superacid catalysts affects the crystal structure, specific surface properties, and surface sulfur species, and that these properties will inevitably cause changes in the acidic properties and catalytic activity of the catalyst. Therefore, in-situ pyridine infrared technology was used to characterize the influence of alumina addition and calcination on catalyst acid content and type. The infrared spectrum and corresponding acid data are shown in Figure 6 and Table 2. It is evident that the addition of alumina to the catalyst will have a particular impact on the acid content and type. According to related literature, the L acid in the sulfated zirconia catalyst comes from the electron withdrawing effect of the high valence state S6+ on the catalyst surface on Zr4+ [27]. As a result, the Zr–O electron cloud shifts and exhibits a specific electron-acquiring ability on the metal Zr. At the same time, the electron-acquiring ability of Zr shifts the hydroxyl group (–OH) of the crystal water on the catalyst, so the crystal water H+ displays B acid acidity. The acidity of the pure ZrO2 catalyst sample is dominated by L acid, the amount of which is approximately twice that of total B acid. Fottinger et al. [28] believed that L acid occupied a dominant position in low-sulfur sulfated zirconia catalysts, which corresponds with the research results. However, this acidic distribution is not conducive to the reaction of brominated alkanes. Several studies have shown that adding an appropriate amount of alumina to superacid catalysts can significantly change the distribution of both L acid and B acid [23,29,30]. Analysis of the composition of the L and B acids in an uncalcined catalyst found it to be similar to pure ZrO2 with added alumina. It is considered that the alumina and ZrO2 in the uncalcined catalyst were not tightly combined, and no Zr–O–Al chemical bond was formed. It can be seen from the pyridine adsorption Fourier-transform infrared (Py–IR) that with the addition of alumina, more B acid centers formed, and the B acid content was about 1.5 times the L acid content. It may be that the introduction of alumina led to the sulfate content increase on the catalyst surface, and since sulfate is a prerequisite for super acidic acid formation, the Zr–O–Al bond enhanced the B acid center.

2.3. Optimization of Reaction Parameters

2.3.1. The Effect of Catalyst Preparation Conditions on the Bromopropane Yield

During the catalyst preparation, it was evident that the preparation conditions had an influence on catalytic effect, and consequently the catalyst preparation conditions were optimized. The calcination temperature, amount of Al added, and acid concentration of the prepared catalyst were carefully considered. Firstly, when the calcination temperature was 450–650 °C, the Zr/Al ratio was 2–6, and the acidification concentration was 0.25–2 mol/L. The catalytic performance of the prepared Al2O3/SO42−/ZrO2 catalyst on the reaction was then assessed. The amount of catalyst was 2.5 wt% of the reaction liquid. As shown in Figure 7, when the reaction conditions were an acid-to-alcohol ratio of 3:1, the reaction temperature and time were 110 °C and 3 h. The higher the calcination temperature, the higher the conversion rate of n-propanol. When the calcination temperature was higher than 600 °C, crystals were formed in the catalyst, and the catalytic performance was highest. When the amount of Zr/Al was increased, the catalyst catalytic performance improved. However, too high an Al content reduced the number of acidic sites in the catalyst, and its performance declined. An increase in acidification concentration obviously increases the catalyst acidity, but an excessively high acidification concentration did not significantly improve the catalyst’s performance. Therefore, after comprehensive screening, a calcination temperature of 600 °C, Zr/Al 4:1 ratio, and acidification concentration of 1 mol/L were determined as appropriate. The catalyst preparation conditions were the most ideal possible in order to optimize the reaction.

2.3.2. The Effect of Reaction Temperature and Reaction Time on Bromopropane Yield

After determining the optimal conditions for preparation of the Al2O3/SO42−/ZrO2 catalyst, the effects of reaction temperature and time on the production of bromopropane via an n-propanol reaction were discussed. With a catalyst dosage of 2.5 wt% of the reaction liquid, a temperature range of 105–125 °C, a controlled reaction time of 3 h, a ratio of 3:1 HBr to n-propanol, the reaction yield was analyzed. As shown in Figure 8, the preparation of bromopropane is essentially a reversible reaction for preparing halogenated alkanes from primary alcohols. It is a bimolecular affinity substitution reaction, so the reaction temperature must not be too high. When the temperature increased, the reaction yield increased slightly. If the temperature was too high, it would not significantly impact the reaction. From the perspective of saving energy, 110 °C was the optimal temperature for the reaction. Additionally, the influence of reaction time on bromopropane yield was analyzed. At 110 °C, with a ratio of HBr to n-propanol of 3:1, the reaction yield was investigated with reaction times between 1–6 h. As shown in the figure, with a 3 h reaction time, the bromopropane yield reached 96.18%. With longer reaction times, the yield did not increase significantly, but after 6 h the conversion rate decreased slightly. This may be due to the increased time allowing a reverse reaction, which in turn affected the yield. Therefore, from the perspectives of cost and energy economy, 110 °C and 3 h were selected as the best reaction and test conditions for the remaining parameter evaluation process.

2.3.3. The Effect of Substrate Concentration and Catalyst Dosage on Bromopropane Yield

It was theorized that a higher bromopropane yield would be obtained under the conditions of a low acid-to-alcohol ratio and low catalyst dosage. Therefore, the influence of the hydrobromic acid and n-propanol ratio, and of the catalyst dosage on reaction yield, was studied. As shown in Figure 9, as the molar ratio of acid to alcohol continued to increase, the bromopropane yield gradually increased. When the molar ratio of hydrobromic acid to n-propanol was 3:1, the bromopropane yield reached its maximum, but as the molar ratio continued to increase, the yield began to decrease. This may be due to the substantial molar ratio causing a decrease in the concentration of n-propanol in the reaction stage, thereby reducing the bromopropane yield. When the amount of Al2O3/SO42−/ZrO2 catalyst was 2.5 wt%, the bromopropane yield reached its peak, and when the amount of catalyst was increased further, the yield decreased. This could be due to the excessive catalyst increasing the number of acidic sites in the solution, and n-propanol dehydrating into ether, which would cause side reactions. Therefore, under optimized conditions, when the ratio of hydrobromic acid to n-propanol was 3:1 and the catalyst amount was 2.5 wt%, the bromopropane yield was greatest.

2.4. Catalytic Stability of Al2O3/SO42−/ZrO2

Reusability is one of the critical characteristics of solid acid catalysts, so the reusability of Al2O3/SO42−/ZrO2 catalysts was explored. The used catalyst was filtered from the reaction solution, washed with deionized water to remove impurities, dried at 110 °C for 24 h, and then calcined for 3 h at 600 °C. As shown in Figure 10, the bromopropane yield could still reach more than 80% after four cycles. FT-IR characterization of the recycled catalyst was carried out, and the results compared with fresh catalyst, as shown in Figure 4B. It was found that the bands at 1240 and 1120 cm−1, which were attributed to the asymmetric O=S=O of the used catalyst, were weakened, so it was considered that the loss of sulfate radicals leads to a reduction in catalytic performance. The addition of Al2O3 dramatically increases the number of acidic sites in the catalyst, so that more acidic sites can be used continuously and simultaneously. This suppresses the decomposition and loss of sulfur-containing species during the reaction process, thereby prolonging the service life of the catalyst. Compared with other types of catalysts and homogeneous catalysts, as shown in Table S1, when Al2O3/SO42−/ZrO2 is used as the catalyst, the TOF at this time is larger and the reaction yield is better. The catalyst was applied to other reactions for preparing brominated alkanes, as shown in Table S2. It was found that the selected catalyst has a good catalytic effect on the bromination of other short-chain alcohols to prepare brominated alkanes. Experiments have found that the catalyst has the same catalytic effect on hindered alcohols. In the experiment of preparing isobromopropane from isopropanol, the use of a catalyst can make the yield of the reaction reach more than 95%.

3. Materials and Methods

3.1. Materials

Zirconium oxychloride (ZrClO2·8H2O) was obtained from Aladdin (Shanghai, China). NH3·H2O, aluminum nitrate (Al(NO3)3·9H2O), N-propanol, and concentrated sulfuric acid were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Hydrobromic acid (40%) was purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Deionized water was self-made by the laboratory. All purchased chemicals were without further purification.

3.2. Catalyst Preparation

In this paper, the co-precipitation method [31,32,33] was used to prepare the catalyst. A certain proportion of ZrClO2·8H2O and Al(NO3)3·9H2O were dissolved in deionized water, and stirring was continued to ensure complete dissolution. Ammonia water was selected as a precipitant and added dropwise to the solution to keep the pH of the slurry at about 8–10; the solution was then left to stand for 24 h. The obtained solid was washed with deionized water until there was no Cl (AgNO3 aqueous solution detection), and dried in an oven at 110 °C for 24 h. The solid was impregnated with a certain concentration of H2SO4 solution for 6 h and then dried in an oven at 110 °C for 24 h. It was then calcined in a muffle furnace for 3 h to obtain the desired catalyst.

3.3. Catalyst Characterization

The XRD spectrum of the sample was measured with an X-ray diffractometer (XRD, Rigaku Ultima IV, Rigaku, Japan), using Cu-Kα radiation (k = 0.154 nm, 40 kV, 40 mA); the scanning rate was 10°/min, and the scanning range was 5–80°.
The XPS spectrum of the sample was measured by Thermo Flyer 250 XI (ThermoFisher K-Alpha, Thermo Fisher Scientific, Shanghai, China) with an energy range of 100–4000 eV. The vacuum degree was 8 × 10−8, and the incident angle was 90°.
The 3H-2000PS1 static volume (Beishide Instrument Technology (Beijing) Co., Ltd., Beijing, China) method was used to analyze the specific surface area and pore size. All samples were degassed in a vacuum at 120 °C for 10 h. The specific surface area was measured by the multipoint Brunauer–Emmett–Teller (BET) method, and the pore size was calculated by the Barrett–Joyner–Halenda (BJH) method.
The FT-IR spectrum of the sample was analyzed on a Nicolet 5700 spectrometer (Thermo Fisher, NYC, Waltham, MA, USA), and the spectral range was from 4000 cm−1 to 400 cm−1. The materials needed to be pretreated with potassium bromide and uniformly ground and compressed.
Pyridine in-situ infrared (Py-IR) was used to detect the Lewis and Brønsted acid sites of the catalyst. The catalyst was pretreated at a pressure of 10−5 Pa and 380 °C for 90 min, then reduced to 80 °C for pyridine adsorption for 30 min, and finally scanned with an infrared spectrometer at 200 °C [34].
In order to observe the microscopic morphology and element distribution of the catalytic material, a field emission scanning electron microscope (SEM, Inspect F50, FEI, Hillsboro, OR, USA) was used.

3.4. Catalytic Performance Research

A certain proportion of hydrobromic acid (HBr, 40%) and n-propanol were added to a round bottom flask. To this, 2 wt% of Al2O3/SO42−/ZrO2 catalyst was added, and the solution was heated and stirred in an oil bath with adjustments to the temperature of the oil bath. A thorn-type rectification column was connected to the flask, and the condensate was collected above the rectification column. After stopping heating at a predefined time, the reaction liquid and condensate were collected for further analysis.

3.5. Product Analysis

The quantitative analysis of bromopropane produced by the reaction of hydrobromic acid and n-propanol was determined by high-performance gas chromatography (GC9160, Shanghai Ouhua) and nuclear magnetism. DMSO is used as a solvent for nuclear magnetic determination. The conversion rate and selectivity of n-propanol are calculated by the following formula:
C o n v e r s i o n = i n i t i a l   m o l a r   a m o u n t   o f   n p r o p a n o l f i n a l   m o l a r   a m o u n t   o f   n p r o p a n o l i n i t i a l   m o l a r   a m o u n t   o f   n p r o p a n o l × 100 % S e l e c t i v i t y = f i n a l   m o l a r   a m o u n t   o f   b r o m o p r o p a n e i n i t i a l   m o l a r   a m o u n t   o f   n p r o p a n o l × 100 %

4. Conclusions

In conclusion, the Al2O3/SO42−/ZrO2 catalyst was successfully prepared by the co-precipitation method. When the calcination temperature was 600 °C, the ratio of Zr/Al was 4:1, and the acidified sulfuric acid concentration was 1 mol/L, the catalyst at this time had the best catalytic performance for the bromination of n-propanol to bromopropane. Under optimal conditions, the ratio of the amount of hydrobromic acid and alcohol was 3:1, and the bromopropane output could reach 96.18% after reacting at 110 °C for 3 h. In a homogeneous reaction, when an ionic liquid and concentrated sulfuric acid are used as catalysts, the reaction temperature is relatively high. At the same time, analysis shows that no by-products are formed, which greatly improves the yield of bromopropane. Catalytic distillation was used to prepare bromopropane, and the product was separated and purified during the reaction, which reduced energy consumption. In addition, the sulfated zirconia catalyst with Al2O3 added could still achieve a yield of more than 80% bromopropane after being recycled 4 times; thus, adding Al2O3 reduced the deactivation rate of pure sulfated zirconia to a certain extent and improved its repetitive use. It was proved that it is feasible to modify sulfated zirconia with Al2O3.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11121464/s1, Figure S1: O42−/MxOy catalytic reaction mechanism, Table S1: TOF of the Al2O3/SO42−/ZrO2 catalyst and comparison with other catalysts, Table S2: Conversion substrate application of the Al2O3/SO42−/ZrO2 catalyst.

Author Contributions

Conceptualization, S.Y. and X.G.; methodology, S.Y.; validation, S.Y.; formal analysis, S.Y.; investigation, S.Y.; resources, S.Y.; data curation, S.Y.; writing—original draft preparation, S.Y.; writing—review and editing, S.Y., X.G., X.P., L.G. (Liuyu Gu) and G.X.; supervision, X.L.; project administration, L.G. (Lijing Gao) and X.P.; funding acquisition, G.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by National Key R&D Program of China (No. 2019YFB1504003).

Conflicts of Interest

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

References

  1. Kiss, A.A.; Dimian, A.C.; Rothenberg, G. Solid acid catalysts for biodiesel production—Towards sustainable energy. Adv. Synth. Catal. 2006, 348, 75–81. [Google Scholar] [CrossRef]
  2. Reddy, B.M.; Patil, M.K. Organic syntheses and transformations catalyzed by sulfated zirconia. Chem. Rev. 2009, 109, 2185–2208. [Google Scholar] [CrossRef] [PubMed]
  3. Yu, G.X.; Zhou, X.L.; Tang, C.; Li, C.L.; Wang, J.A.; Novaro, O. A comparative study of the synthesis approaches and catalytic behaviors of Pt/SO42−/ZrO2-Al2O3 catalysts for n-hexane hydro isomerization. Catal. Commun. 2008, 9, 1770–1774. [Google Scholar] [CrossRef]
  4. Matsuhashi, H.; Hino, M.; Arata, K. Synthesis of the solid superacid catalyst of tin oxide treated with sulfate ion. Appl. Catal. 1990, 59, 205–212. [Google Scholar] [CrossRef]
  5. Hino, M.; Arata, K. Synthesis of highly active superacids of SO4/ZrO2 with Ir, Pt, Rh, Ru, Os, and Pd substances for reaction of butane. Catal. Lett. 1994, 30, 25–30. [Google Scholar] [CrossRef]
  6. Song, H.; Wang, N.; Song, H.L.; Li, F. La–Ni modified S2O82−/ZrO2-Al2O3 catalyst in n-pentane hydroisomerization. Catal. Commun. 2015, 59, 61–64. [Google Scholar] [CrossRef]
  7. Canton, P.; Olindo, R.; Pinna, F.; Strukul, G.; Riello, P.; Meneghetti, M.; Cerrato, G.; Morterra, C.; Benedetti, A. Alumina-promoted sulfated zirconia system: Structure and microstructure characterization. Chem. Mater. 2001, 13, 1634–1641. [Google Scholar] [CrossRef]
  8. Song, H.; Wang, N.; Song, H.L.; Li, F.; Jin, Z.S. Effect of Al content on the isomerization performance of solid superacid Pd-S2O82-/ZrO2-Al2O3. Chin. J. Chem. Eng. 2014, 22, 1226–1231. [Google Scholar] [CrossRef]
  9. Kamoun, N.; Younes, M.K.; Ghorbel, A.; Mamede, A.S.; Rives, A. Comparative study of aerogels nanostructured catalysts: Ni/ZrO2-SO42− and Ni/ZrO2-Al2O3-SO42−. Ionics 2015, 21, 221–229. [Google Scholar] [CrossRef]
  10. Kim, S.Y.; Lohitharn, N.; Goodwin, J.G., Jr.; Olindo, R.; Pinna, F.; Canton, P. The effect of Al2O3-promotion of sulfated zirconia on n-butane isomerization: An isotopic transient kinetic analysis. Catal. Commun. 2006, 7, 209–213. [Google Scholar] [CrossRef]
  11. Cao, C.J.; Han, S.; Chen, C.L.; Xu, N.P.; Mou, C.Y. Hydroisomerization of n-hexane over gallium-promoted sulfated zirconia. Catal. Commun. 2003, 4, 511–515. [Google Scholar] [CrossRef]
  12. Cao, C.J.; Yu, X.Z.; Chen, C.L.; Xu, N.P.; Wang, Y.R.; Mou, C.Y. Comparison of the promoting effects of gallium and aluminum on the n-butane isomerization activity of sulfated zirconia. React. Kinet. Catal. Lett. 2004, 83, 85–92. [Google Scholar] [CrossRef]
  13. Atlas, C. Search for anomalous production of prompt same-sign lepton pairs and pair-produced doubly charged Higgs bosons with √s = 8 TeV pp collisions using the ATLAS detector. J. High Energy Phys. 2015, 3, 41. [Google Scholar]
  14. Kanemura, S.; Yagyu, K.; Yokoya, H. First constraint on the mass of doubly charged Higgs bosons in the same-sign diboson decay at the LHC. Phys. Lett. 2013, 726, 316–319. [Google Scholar] [CrossRef] [Green Version]
  15. Li, X.B.; Nagaoka, K.; Lercher, J.A. Labile sulfates as key components in active sulfated zirconia for n-butane isomerization at low temperatures. J. Catal. 2004, 227, 130–137. [Google Scholar] [CrossRef]
  16. Ma, Z.M.; Meng, X.; Liu, N.W.; Yang, C.; Shi, L. Preparation, characterization, and isomerization catalytic performance of palladium loaded zirconium hydroxide/sulfated zirconia. Ind. Eng. Chem. Res. 2018, 57, 14377–14385. [Google Scholar] [CrossRef]
  17. Zalewski, D.J.; Alerasool, S.; Doolin, P.K. Characterization of catalytically active sulfated zirconia. Catal. Today 1999, 53, 419–432. [Google Scholar] [CrossRef]
  18. Mercera, P.D.L.; Ommen, J.G.V.; Doesburg, E.B.M.; Burggraaf, A.J.; Boss, J.R.H. Zirconia as a support for catalysts: Evolution of the texture and structure on calcination in air. Appl. Catal. 1990, 57, 127–148. [Google Scholar] [CrossRef] [Green Version]
  19. Wang, P.; Yang, S.W.; Kondo, J.N.; Domen, K.; Yamada, T.; Hattori, H. Spectroscopic study of H2 and CO adsorption on platinum-promoted sulfated zirconia catalysts. J. Phys. Chem. B 2003, 107, 11951–11959. [Google Scholar] [CrossRef]
  20. Chen, W.H.; Ko, H.H.; Sakthivel, A.; Huang, S.J.; Liu, S.H.; Lo, A.Y.; Tsai, T.C.; Liu, S.B. A solid-state NMR, FT-IR and TPD study on acid properties of sulfated and metal-promoted zirconia: Influence of promoter and sulfation treatment. Catal. Today 2006, 116, 111–120. [Google Scholar] [CrossRef]
  21. Wang, Z.C.; Shui, H.F.; Zhang, D.X.; Gao, J.S. A comparison of FeS, FeS + S and solid superacid catalytic properties for coal hydro-liquefaction. Fuel 2007, 86, 835–842. [Google Scholar] [CrossRef]
  22. Cavani, F.; Guidetti, S.; Trevisanut, C.; Ghedini, E.; Signoretto, M. Unexpected events in sulfated zirconia catalyst during glycerol-to-acrolein conversion. Appl. Catal. A Gen. 2011, 409, 267–278. [Google Scholar] [CrossRef]
  23. Song, H.; Wang, N.; Song, H.L.; Li, F. Effect of Pd content on the isomerization performance over Pd-S2O82−/ZrO2-Al2O3 catalyst. Res. Chem. Intermed. 2016, 42, 951–962. [Google Scholar] [CrossRef]
  24. Vishwanathan, V.; Balakrishna, G.; Rajesh, B.; Jayasri, V.; Sikhwivhilu, L.M.; Coville, N.J. Alkylation of catechol with methanol to give guaiacol over sulphate-modified zirconia solid acid catalysts: The influence of structural modification of zirconia on catalytic performance. Catal. Commun. 2008, 9, 2422–2427. [Google Scholar] [CrossRef]
  25. Appay, M.D.; Manoli, J.M.; Potvin, C.; Muhler, M.; Wild, U.; Pozdnyakova, O.; Paal, Z. High-resolution electron microscopic, spectroscopic, and catalytic studies of intentionally sulfided Pt/ZrO2–SO4 catalysts. J. Catal. 2005, 222, 419–428. [Google Scholar] [CrossRef]
  26. Breitkopf, C.; Matysik, S.; Papp, H. Selective poisoning of active centers of sulfated zirconia monitored by TAP, XPS, and DRIFTS. Appl. Catal. A Gen. 2006, 301, 1–8. [Google Scholar] [CrossRef]
  27. Comelli, R.A.; Vera, C.R.; Parera, J.M. Influence of ZrO2 crystalline structure and sulfate Ion concentration on the catalytic activity of SO42−-ZrO2. J. Catal. 1995, 151, 96–101. [Google Scholar] [CrossRef]
  28. Fottinger, K.; Zorn, K.; Vinek, H. Influence of the sulfate content on the activity of Pt containing sulfated zirconia. Appl. Catal. A Gen. 2005, 284, 69–75. [Google Scholar] [CrossRef]
  29. Barrera, A.; Montoya, J.A.; Viniegra, M.; Navarrete, J.; Espinosa, G.; Vargas, A.; Angel, P.D.; Perez, G. Isomerization of n-hexane over mono-and bimetallic Pd–Pt catalysts supported on ZrO2–Al2O3–WOx prepared by sol–gel. Appl. Catal. A Gen. 2005, 290, 97–109. [Google Scholar] [CrossRef]
  30. Zhukov, Y.M.; Efimov, A.Y.; Shelyapina, M.G.; Petranovskii, V.; Zhizhin, E.V.; Burovikhina, A.; Zvereva, I.A. Effect of preparation method on the valence state and encirclement of copper exchange ions in mordenites. Microporous Mesoporous Mater. 2016, 224, 415–419. [Google Scholar] [CrossRef]
  31. Hua, W.M.; Xia, Y.D.; Yue, Y.H.; Gao, Z. Promoting effect of Al on SO42−/MxOy (M = Zr, Ti, Fe) catalysts. J. Catal. 2000, 196, 104–114. [Google Scholar] [CrossRef]
  32. Wang, J.H.; Mou, C.Y. Alumina-promoted mesoporous sulfated zirconia: A catalyst for n-butane isomerization. Appl. Catal. A Gen. 2005, 286, 128–136. [Google Scholar] [CrossRef]
  33. Yang, Y.C.; Weng, H.S. Al-promoted Pt/SO42−/ZrO2 with low sulfate content for n-heptane isomerization. Appl. Catal. A. Gen. 2010, 384, 94–100. [Google Scholar] [CrossRef]
  34. Taouli, A.; Klemt, A.; Breede, M.; Reschtilowski, W. Acidity investigations and determination of integrated molar extinction coefficients for infrared absorption bands of ammonia adsorbed on acids sites of MCM-41. Stud. Surf. Sci. Catal. 1999, 125, 307–314. [Google Scholar]
Figure 1. XRD patterns of Al2O3/SO42−/ZrO2 catalysts with different calcination temperatures (A) and different loadings (B).
Figure 1. XRD patterns of Al2O3/SO42−/ZrO2 catalysts with different calcination temperatures (A) and different loadings (B).
Catalysts 11 01464 g001
Figure 2. SEM images of catalyst before (A) and after (B) modification.
Figure 2. SEM images of catalyst before (A) and after (B) modification.
Catalysts 11 01464 g002
Figure 3. EDS images of prepared Al2O3/SO42−/ZrO2 catalysts.
Figure 3. EDS images of prepared Al2O3/SO42−/ZrO2 catalysts.
Catalysts 11 01464 g003
Figure 4. (A) FT–IR spectra of Al2O3/SO42−/ZrO2 catalysts with different loadings; (B) FT–IR spectra of Al2O3/SO42−/ZrO2 catalysts before and after the fourth circle.
Figure 4. (A) FT–IR spectra of Al2O3/SO42−/ZrO2 catalysts with different loadings; (B) FT–IR spectra of Al2O3/SO42−/ZrO2 catalysts before and after the fourth circle.
Catalysts 11 01464 g004
Figure 5. (A) XPS spectra of Al2O3/SO42−/ZrO2 catalysts; (B) XPS spectra of S 2p for Al2O3/SO42−/ZrO2 catalysts.
Figure 5. (A) XPS spectra of Al2O3/SO42−/ZrO2 catalysts; (B) XPS spectra of S 2p for Al2O3/SO42−/ZrO2 catalysts.
Catalysts 11 01464 g005
Figure 6. Py-IR spectra of prepared SO42−/ZrO2 and Al2O3/SO42−/ZrO2 catalysts.
Figure 6. Py-IR spectra of prepared SO42−/ZrO2 and Al2O3/SO42−/ZrO2 catalysts.
Catalysts 11 01464 g006
Figure 7. Bromopropane yield at different calcination temperatures (A), Zr/Al ratio (B), and acid concentration (C). Reaction conditions: raw material ratio: 3:1, 110 °C, 3 h.
Figure 7. Bromopropane yield at different calcination temperatures (A), Zr/Al ratio (B), and acid concentration (C). Reaction conditions: raw material ratio: 3:1, 110 °C, 3 h.
Catalysts 11 01464 g007
Figure 8. Bromopropane yield at different temperatures (A) and times (B). Reaction conditions: catalyst (2.5 wt%), material ratio: 3:1.
Figure 8. Bromopropane yield at different temperatures (A) and times (B). Reaction conditions: catalyst (2.5 wt%), material ratio: 3:1.
Catalysts 11 01464 g008
Figure 9. Bromopropane yield at different molar ratios (A) and catalyst dosages (B). Reaction conditions: 110 °C, 3 h.
Figure 9. Bromopropane yield at different molar ratios (A) and catalyst dosages (B). Reaction conditions: 110 °C, 3 h.
Catalysts 11 01464 g009
Figure 10. The recycled Al2O3/SO42−/ZrO2 catalyst for bromopropane yield. Reaction conditions: raw material ratio: 3:1, 110 °C, 3 h.
Figure 10. The recycled Al2O3/SO42−/ZrO2 catalyst for bromopropane yield. Reaction conditions: raw material ratio: 3:1, 110 °C, 3 h.
Catalysts 11 01464 g010
Table 1. Comparison of surface area and pore size for prepared SO42−/ZrO2 and Al2O3/ SO42−/ZrO2.
Table 1. Comparison of surface area and pore size for prepared SO42−/ZrO2 and Al2O3/ SO42−/ZrO2.
CatalystsSBET a (m2/g)Dmean b (nm)
ZrO243.550.27
2(Zr/Al)69.633.65
3(Zr/Al)57.837.60
4(Zr/Al)105.926.80
5(Zr/Al)98.122.23
6(Zr/Al)87.98.98
a BET surface area was determined from N2 adsorption isotherm. b Average pore size was determined from the desorption average pore diameter (4 V/A by BET).
Table 2. Acidity of prepared SO42−/ZrO2 and Al2O3/ SO42−/ZrO2 catalysts.
Table 2. Acidity of prepared SO42−/ZrO2 and Al2O3/ SO42−/ZrO2 catalysts.
CatalysisBrønsted Acid Sites (µmol/g)Lewis Acid Sites (µmol/g)Total Acids (µmol/g)Ratio of Brønsted
to Lewis
ZrO221.7739.9061.670.55
(4)Zr/Al Uncalcined12.6922.2724.960.57
(4)Zr/Al8.695.6214.311.54
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, S.; Guo, X.; Pan, X.; Gu, L.; Liu, X.; Gao, L.; Xiao, G. Synthesis of Brominated Alkanes via Heterogeneous Catalytic Distillation over Al2O3/SO42−/ZrO2. Catalysts 2021, 11, 1464. https://doi.org/10.3390/catal11121464

AMA Style

Yang S, Guo X, Pan X, Gu L, Liu X, Gao L, Xiao G. Synthesis of Brominated Alkanes via Heterogeneous Catalytic Distillation over Al2O3/SO42−/ZrO2. Catalysts. 2021; 11(12):1464. https://doi.org/10.3390/catal11121464

Chicago/Turabian Style

Yang, Su, Xiaoxuan Guo, Xiaomei Pan, Liuyu Gu, Xueping Liu, Lijing Gao, and Guomin Xiao. 2021. "Synthesis of Brominated Alkanes via Heterogeneous Catalytic Distillation over Al2O3/SO42−/ZrO2" Catalysts 11, no. 12: 1464. https://doi.org/10.3390/catal11121464

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