Open access peer-reviewed chapter

Environmental-Friendly Catalytic Oxidation Processes Based on Hierarchical Titanium Silicate Zeolites at SINOPEC

Written By

Changjiu Xia, Xinxin Peng, Yao Zhang, Baorong Wang, Min Lin, Bin Zhu, Yibin Luo and Xingtian Shu

Submitted: 08 November 2016 Reviewed: 08 March 2017 Published: 05 July 2017

DOI: 10.5772/intechopen.68389

From the Edited Volume

Green Chemical Processing and Synthesis

Edited by Iyad Karame and Hassan Srour

Chapter metrics overview

2,736 Chapter Downloads

View Full Metrics

Abstract

Since it was claimed by EniChem in 1983 for the first time, titanium silicate‐1 (TS‐1) zeolite presented the most delightful catalytic performance in the area of selective organic oxidation reactions. To enhance the mass diffusion property, hierarchical titanium silicate with hollow cavities within crystal was prepared by using a post‐synthesis treatment in the presence of organic template, and then, it was commercially produced and employed in many industrial catalytic oxidation processes, such as propylene epoxidation, phenol hydroxylation, and cyclohexanone ammoximation. Moreover, we also developed several totally novel oxidation reactions on hollow titanium silicate (HTS) zeolite, i.e., Baeyer‐Villiger oxidation of cyclohexanone and chlorohydrination of allyl chloride with HCl and H2O2. In all cases, HTS shows much better catalytic performance than TS‐1, attributing to the mass diffusion intensification by introducing hollow cavities. On the other hand, enormous works on synthesizing hierarchical TS‐1 zeolites with open intracrystalline mesopores have been done via silanization treatment and recrystallization. Based on them, several bulk molecule oxidation processes with tert‐butyl hydroperoxide, such as epoxidation of fatty acid methyl ester (FAME) and large olefins, have been carried out. As a consequence, hierarchical TS‐1 zeolites supply a platform for developing environmental‐friendly catalytic oxidation processes to remarkably overcome the drawbacks of traditional routes.

Keywords

  • titanium silicate zeolite
  • catalytic oxidation
  • novel chlorohydrination process
  • HPPO
  • hierarchical structure
  • hydrogen peroxide

1. Introduction

Framework‐incorporated zeolite is an important kind of acid catalyst, which has already been successfully applied or shown great application potentials in many industrial processes in last 30 years [14]. Since the first claim of titanium silicate‐1 (TS‐1) zeolite by EniChem in 1983, serial of metal elements (such as Ti, Sn, Fe, V, Cr, Mn, Zr, Ga, and Ge) have been incorporated into the framework matrix of zeotype materials via the isomorphous substitution of Al atoms in crystalline aluminosilicates through different approaches [510]. When Ti4+, Sn4+, and Zr4+ species incorporated into the framework of zeolites, basic subunits are neutral, without needing balancing by cation ions, thus no formation of Brönsted acid sites. However, these metal ions can accept shared electron pairs from electron‐rich groups (i.e., peroxide, carbonyl, and C-C bonds) to make substrates much more reactive during catalytic reactions [1113]. On the other hand, catalytic selective oxidation of hydrocarbons to produce oxygenated organic compounds is of great significance to give birth to chemical intermediates and fuels. Unfortunately, traditional oxidation approaches always suffer from using hazard reagents, complex operation, high energy and cost consumption, low product selectivity, and high pollution [1416].

To avoid these drawbacks, heterogeneous catalysts (specially based on zeolites) have been introduced to develop highly efficient oxygenated hydrocarbons transformation due to their unique porosity and Lewis acidity for activating oxidants and/or subtracts [1719]. For example, in conventional route, cyclohexanone oxime is produced via a hydroxylamine method, which needs four steps, involving ammonia oxidation, production of hydroxylamine with liquid acids, cyclohexanone ammoximation, and decomposition of ammonium, generating a large amount of waste and side products [2022]. In 1986, a totally novel “one‐step” cyclohexanone ammoximation process with ammonia (NH3) and hydrogen peroxide (H2O2) was developed by EniChem via using TS‐1 zeolite as a catalyst under mild conditions, with water only as a by‐product [23, 24]. Obviously, Ti‐incorporated zeolites provide excellent opportunity to update the efficiency of oxidation processes. However, TS‐1 zeolite suffered the limitations of mass diffusion, its narrow 10 member‐ring channel [2527]. Thus, a novel post‐synthesis route to modify the traditional TS‐1 was developed by the researchers at SINOPEC before 2000, obtaining the far more stable hollow titanium silicate (HTS) zeolite. It is worthy to note that HTS zeolite has been commercially manufactured by SINOPEC, and applied in several green catalytic oxidation processes, such as cyclohexanone ammoximation, phenol hydroxylation, and propylene epoxidation [2830]. Moreover, several completely new reactions, such as chlorohydrination of allyl chloride and Baeyer‐Villiger oxidation of cyclohexanone, were initially developed on using HTS zeolite [31, 32]. In recent years, a spongy‐like TS‐1 zeolite with abundant of open mesopores was prepared via combining the silanization treatment and post‐synthesis route under hydrothermal conditions. And it also found some bulk molecule oxidation reactions based on this hierarchical TS‐1, such as epoxidation of methyl oleate and large olefins. Thus, this review will majorly summarize the application of hierarchical TS‐1 we synthesized in the area of oxidation, as shown in Figure 1.

Figure 1.

Catalytic oxidation reactions on hierarchical titanium silicate zeolite with H2O2 as an oxidant under mild conditions.

Advertisement

2. Synthesis strategies of hierarchical titanium silicate zeolites

TS‐1 zeolite was claimed by the researcher EniChem in 1983 for the first time, and then it attracted kind attentions from the world, owing to its pioneering role in exploiting metal‐incorporated zeolites [4]. However, in the early years, the key factors for TS‐1 zeolite synthesis were far beyond well understanding, thus it was hard to repeat the synthesis of TS‐1 in parallel, even under the same conditions. That is ascribed to the different crystallization behaviors between Ti and Si precursors during hydrothermal synthesis processes. To overcome this major drawback, a post‐synthesis method, involving the simultaneous dissolution of framework atoms and recrystallization of low aggregated species, was developed by M. Lin and coworkers at SINOPEC in the end of 1990s [3335]. In a typical procedure, TS‐1 was synthesized via the EniChem published method for the first time; and then it was calcined and treated by specific agents; in the end, TS‐1 was recrystallized in the presence of SODs at high temperature with autogenous pressure. It has been confirmed that unique intracrystalline cavities (about several to several tens nm in size) are generated through post‐treatment, and this material is referred to as hollow titanium silicate (HTS) zeolite, which is of much better diffusion property than TS‐1, as shown in Figure 2 [36]. On the other hand, recrystallization treatment promotes the increasing of framework Ti content, with the significant decreasing of framework defects (Si‐OH and Ti‐OH groups). Thus, HTS displays far higher stability than TS‐1 during catalytic oxidation processes, especially in alkaline ammoximation reactions, due to high resistance to ammonia. Up to now, HTS zeolite had been already commercially manufactured at SINOPEC, and successfully applied in many selective catalytic oxidation processes, i.e., cyclohexanone ammoximation, propylene epoxidation, and phenol hydroxylation, presenting desirable economic and social benefits via reducing cost and waste production. To the best of our best knowledge, HTS is the first and only hierarchical titanium‐based zeolite applied in industry, and the detailed synthesis, characterization, and application have been summarized in our previous chapter [37].

Figure 2.

Transmission electron microscopy images of TS‐1 and hierarchical TS‐1 zeolites : (A) TS‐1, (B) HTS, (C) TS‐OS, and (D) TS‐SP.

However, these cavities inside HTS zeolite cannot favor the accessibility of bulk organic molecules to inner tetrahedral framework Ti sites, owing to the 10‐member ring pore mouth. To solve this problem, many attempts and efforts have been paid through introducing mesopores and/or macropores with zeolite particles, including selective removing framework‐composite atoms, channel filling with soft or hard templates, reassembly of small crystals around mesopore templates, and adding amphiphilic organic surfactants and silicon sources. Among them, hierarchical TS‐1 zeolite of rich open mesopores (remarked as TS‐OS, as shown in Figure 2C) synthesized by introducing amphiphilic organic silane agents is considered as another potential industrial catalyst in the near future [38, 39]. It is very easy and available to synthesis TS‐OS even in industry, just adding a few amount of organic silane agents, such as N‐[3‐(trimethoxysilyl) propyl]aniline and N,N‐diethyl‐3‐(trimethoxysilyl) propan‐1‐amine, into conventional TS‐1 synthesis mixture under hydrothermal conditions. That is attributed to the inhibition of crystal growth by organic silane groups inside particles, as Si‐C bonds are ultrastable at high temperature. It is noted that silane groups are easily aggregated, thus mesopores are formed when the organic groups are removed by calcination in air at 550°C. Of course, compared with TS‐1, TS‐OS is of high mass transfer capability, allowing the entrance of bulk organic subtracts and oxidants. Thus, TS‐OS exhibits much better catalytic performance than TS‐1 in some specific oxidations. It is worthy to mention that organic silane, usually serving as coupling agents for binder and coating, are very cheap and easily obtained in market, so it is available to synthesis of TS‐OS at large scale.

In addition, to make inner mesopores be more uniform, TS‐OS zeolite was post‐synthesized in the presence of long chain quaternary ammonium hydroxide (such as cetyltrimethylammonium hydroxide, CTAOH) under hydrothermal conditions, to give birth to spongy‐like TS‐1 zeolite (labeled as TS‐SP), as shown in Figure 2D. As CTAOH can enter into the crystal of TS‐OS along mesopores rather than micropores, the dissolution of framework atoms and reassembly of soluble Ti and Si species under the impact of CTAOH micelles with zeolite particles are simultaneously promoted during post‐synthesis processing, thus forming the relatively uniform mesopores, as demonstrated in Figure 3. There are two peaks existed for TS‐OS, the first one is ascribed to mesopores created by silane agents and the second one is related to intercrystalline mesopore among small crystal particles. The band for TS‐SP is very narrow, suggesting the intracrystalline mesopores become much more uniform after reassembly.

Figure 3.

Pore‐size distribution of TS‐OS and TS‐SP zeolites calculated by the nonlocal density function theory (NL‐DFT) method.

Advertisement

3. Nature of the activating of H2O2 molecules by TS‐1 zeolite

As it is well known that TS‐1 zeolite is an excellent catalyst via improving the nucleophilic attack capability of aqueous H2O2 solution in catalytic oxidation reactions [40, 41]. When the H2O2 molecules are adsorbed on the tetrahedral Ti active sites (Figure 4), the negative charges on both Oa and Ob atoms in TS2 are −0.460 and −0.476, respectively, which are of much stronger nucleophilic attack capability than that in H2O2 molecules (about −0.426), as listed in Table 1. It indicates that the H2O2 molecules are activated, with the formation of Ti‐OOH species.

Figure 4.

Activation of H2O2 molecules by tetrahedral framework Ti sites calculated by the density function theory (DFT) method.

Atoms Before H2O2 adsorption After H2O2 adsorption Charge difference
H2O2 Oa −0.426 −0.460 −0.034
Ob −0.426 −0.476 −0.050
Ha 0.426 0.410 −0.016
Hb 0.426 0.469 0.043

Table 1.

Charge difference of TS‐1 cluster and H2O2 molecule before and after H2O2 adsorption.

Advertisement

4. Chlorohydrination of allyl chloride with HCl and H2O2 to produce dichloropropanol

Recently, we have developed a totally novel route to prepare dichloropropanol (DCP) through the chlorohydrination of allyl chloride (AC) with HCl and H2O2 base on HTS as a catalyst, to replaced Cl2 and H2O as raw materials in traditional route, as shown in Scheme 1 [42]; that is, because Cl2 is highly hazard and harmful for environment and human health, and difficult to be stored and transfer, producing a large amount of waste. Most importantly, to avoid the production of trichloropropane (TCP), a great deal of water are added to reduce the concentration of Cl2 and its derivatives (HOCl and HCl), therefore causing the low content of DCP (<4 wt%) in final mixture [43, 44]. Obviously, the tradition Cl2 route goes against the principles of green and sustainable chemistry, but HCl is very available and less harmful. It is surprisingly to observe that both AC conversion and DCP selectivity simultaneously over 95% under optimized reaction conditions, and DCP concentration is higher than 20 wt% in aqueous solution. Thus, this novel route demonstrates very excellent industrial potential, due to its high efficiency and low waste produced.

Scheme 1.

Comparison of traditional and novel chlorohydrination reactions.

To study the mechanism of this route, different Ti‐containing catalysts were evaluated by a novel chlorohydrination reaction under almost the same condition, as shown in Figure 5. Among the catalysts we tested, HTS shows the highest AC conversion (over 85%), far higher than that of TS‐1 (49%), owing to the improvement of mass diffusion through hierarchically closed mesopores within HTS zeolite particles. At the same time, HTS and TS‐1 exhibit much higher selectivity (>88%) of DCP (including 1,3‐DCP and 2,3‐DCP) than other catalysts, suggesting tetrahedral framework Ti species help the formation of DCP. And the major product catalyzed by TS‐1 and HTS is 1,3‐DCP, while 2,3‐DCP is mainly obtained via the traditional Cl2 route. To explicate this phenomena, ring‐opening reaction of epichlorohydrin (ECH) with HCl in aqueous solution had been carried out, with a 1,3‐DCP selectivity higher than 98% and complete AC conversion. In fact, TS‐1 zeolite is of high catalytic activity in olefin epoxidation reaction, and many H2O2 oxidized propylene epoxidation units (called HPPO process) have been established in global. However, TS‐1 shows poor catalytic performance in the epoxidation of AC with H2O2 due to the strong electron‐withdrawing effect of Cl atoms on C=C bonds [45, 46]. However, in the presence of HCl, the epoxidation reaction can be accelerated, which is ascribed to the promotion of chemical balance to DCP via ring‐opening reaction catalyzed by H+ species.

Figure 5.

Catalytic performance of different Ti‐based catalysts in novel chlorohydrination of AC with HCl and H2O2 under mild conditions: (A) AC conversion and (B) DCP selectivity.

It is worthy to note that there are still many serious drawbacks related to the HPPO process, attributing to the side‐products caused by H2O [47, 48]. As a result, this route provides a novel method to prepare DCP, with high AC conversion and target product selectivity. Above all, it is indicated that the mechanism of this route lines in the first epoxidation of AC catalyzed by framework Ti species, and then ECH becomes ring opening with HCl. On the other hand, Cl2 and its derivatives (such as Cl3, but no HOCl) are detected by UV‐Raman spectroscopy (with an irradiation wavelength of 325 nm). It means that the Cl2‐based chlorohydrination process can also be occurred in this route, similar as the traditional route, inferring that epoxidation of AC competes with the route of Cl2 [4952]. What is more, there is only TCP generated catalyzed by HTS and TS‐1 zeolites, which is relevant with the Cl2 formed via oxidation of HCl with H2O2, significantly increasing the utilization efficiency of AC to prepare DCP. Except for the desired catalytic activity and selectivity, HTS also gives much higher stability than TS‐1 when they are treated by 37 wt% HCl solution at room temperature. It is found that, after 3 months treatment, HTS zeolite still remains as high AC conversion as fresh one, while the catalytic performance of TS‐1 becomes fast decreasing when it was acidic treated for 5 days. As shown in Figure 6, HTS can be used six times in a novel chlorohydrination route, without any decreasing of AC conversion and DCP selectivity, which is attributed to the reincorporation of Ti species into framework through post‐synthesis treatment.

Figure 6.

Conversion of AC and DCP selectivity in the reuse experiments six times under the same reaction conditions.

Advertisement

5. Ammoximation of ketones with H2O2 and NH3

Ammoximation reaction is very important to produce oximes from ketones and aldehydes in industry [5355]. For example, cyclohexanone oxime, which is a key intermediate for manufacturing ε‐caprolactam, is conventionally prepared via the hydroxylamine oxidation route, including hydroxylamine sulfate oxidation (HSO) and hydroxylamine phosphoric oxidation (HPO) approaches [56, 57]. Generally, it involves many harsh drawbacks in these approaches, i.e., long and complex reaction pathways, high energy and material consumed, generating a large amount of waste and side‐products, and harmful for health and environment, as shown in Scheme 2. To overcome these drawbacks, a “one‐step” cyclohexanone ammoximation process with H2O2 and NH3 was developed by EniChem in 1986 using TS‐1 as a catalyst under very moderate conditions [58, 59]. Compared with traditional route, the novel one is highly clean and environmental friendly, with evidently reducing of waste production and capital cost, only water as a side product. The first 12 Kt/a demonstration unit with TS‐1 catalyst had been built up by EniChem in 1994, and a commercial 60 Kt/a unit was carried out by Sumitomo in 2003. By using HTS as a catalyst, a 140 Kt/a commercial cyclohexanone ammoximation process was built by SINOPEC, and it run smoothly for over 10 years since 2000 [60]. Similar to this process, acetone ammoximation with H2O2 and NH3 catalyzed by HTS also carried out, with high acetone conversion and oxime selectivity [61].

Scheme 2.

Comparison of traditional HSO (HPO) route and novel TS‐1 catalyzed ammoximation reaction for the production of cyclohexanone oxime.

The mechanism of “one‐pot” cyclohexanone ammoximation bases on the in‐situ formation of hydroxylamine (NH2OH) through the oxidation of NH3 with H2O2 under the effect of tetrahedral Ti species, and then it reacts with cyclohexanone to form oxime [62, 63]. The coordination of NH3 and H2O2 with framework Ti sites has been investigated by a spectroscopic method, perfectly with an octahedral coordination. And the cyclohexanone ammoximation reaction was operated in a stirred reactor through a continuous liquid phase flow, with a slurry medium (TS‐1 catalyst concentration is about 2–3 wt%) at 80–90°C. It is noted that the TS‐1 and HTS zeolite particles are about 300–500 nm in size, which are too small to be separated by conventional methods and equipment. Thus, the separation of catalyst from reaction mixture and solvent (tert‐butanol) is usually carried out in a membrane filter, and then crude oxime is withdrawn from the bottom of this reactor for further purification, and the spent catalyst should be separated from this reaction system and calcined to remove the organic deposit, which fills the channel of TS‐1 zeolite, as shown in Figure 7.

Figure 7.

Diagrammatic sketch of a “one‐step” cyclohexanone ammoximation process catalyzed by TS‐1 zeolite under moderate conditions.

As shown in Figure 8, we can see that HTS zeolite is of far higher catalytic performance and longer lifetime than TS‐1 under the same conditions. This is because hollow cavities well enhance the mass diffusion within zeolite particles, and relative short residence time of cyclohexanone favors the decrease of side products, especially for heavy organic deposits. Moreover, the ultrahigh stability of HTS is attributed to the recrystallization of zeolite defects through dehydrated condensation between Si‐OH and Ti‐OH groups, making it more tolerant to alkaline intermediates.

Figure 8.

Comparison of catalytic performance of HTS and TS‐1 zeolites in the cyclohexanone ammoximation process with H2O2 and NH3.

In addition, in our previous works, the irreversible deactivation of HTS zeolite has been investigated by using multiple characterization methods and catalytic evaluations [6466]. It is confirmed that the dissolution of framework Si atoms and the removal of framework Ti atoms cause the formation of acidic amorphous TiO2‐SiO2 nanoparticles, leading to the inefficient decomposition of H2O2 molecules during catalytic oxidation processes. At the same time, the simulated deactivation of HTS in aqueous NH3 solution was also examined, and finally, an efficient and easy regeneration method has been developed.

Advertisement

6. Epoxidation of olefins with H2O2

Olefin epoxidation with H2O2 to prepare epoxy compound is an ultrakey reaction for manufacturing chemical intermediates, which is highly relevant with all aspects of human life and social production [6769]. Particularly, propylene oxide (PO), one of the three large propylene derivatives, is majorly produced via a combined chlorohydrination‐saponification technology and a cooxidation method in traditional route, as shown in Scheme 3. Chlorohydrination route lines in the oxidative addition of propylene with Cl2 and H2O to give birth to chloropropanol and HCl, and then chloropropanol reacts with Ca(OH)2, with the formation of propylene oxide. We can conclude that the O atom in PO molecule comes from the H2O molecule, and a large amount of hazardous by‐products and wastewater are caused by Cl2 and Ca(OH)2 [7072]. As a result, a much greener PO manufacture technology catalyzed on TS‐1 (remarked as HPPO) was exploited by EniChem, and then several commercial HPPO units based on TS‐1catalysts were built up and run smoothly in around 2010, significantly reducing the generation of wastewater and unwanted products [73, 74].

Scheme 3.

Reaction networks of chlorohydrination route and cooxidation route for the production of propylene oxide in industry.

Scheme 4 presents the reaction pathways of epoxidation of propylene with H2O2 under the impact of tetrahedral Ti species, while the five‐member ring (originated from framework Ti site, solvent, and H2O2) mechanism is well accepted, as shown in Scheme 4. To verify this mechanism, many characterization and calculation methods were employed and summarized. Since 2000, HTS zeolite was also applied in the HPPO process, showing extreme high catalytic performance and stability. Moreover, the commercial 100 kt/a HPPO unit had been established and run well for over 6000 h, based on the promising result of pilot unit (at a 10 kt/a scale), as illustrated in Figure 9. The demonstration process of HPPO catalyzed by HTS was carried out in a fixed‐bed reactor under mild conditions, with over 97% H2O2 conversion and 96–99% PO selectivity at the same time, as shown in Figure 10. It is even higher than that reported by BASF and DOW (the selectivity of PO is about 93.2%), attributing to the benefits of intracrystalline mesopores [75].

Scheme 4.

Reaction mechanism of propylene epoxidation with H2O2 catalyzed by TS‐1 zeolite.

Figure 9.

Diagrammatic sketch of propylene epoxidation with H2O2 catalyzed by HTS zeolite.

Figure 10.

Catalytic performance of HTS zeolite in propylene epoxidation at pilot scale.

In addition, after the purification of crude PO in a distillation column, the final PO product has a high purity of 99.97%, which had already been used as chemical intermediates to manufacture 1,2‐ethyl glycerol, oilfield demulsifier, propylene carbonate, triisopropanolamine, hydroxypropyl, dimethyl carbonate in market. Up to date, one even larger scaled (300 kt/a) industrial HPPO demonstration unit is under constructing, based on the previous experiment results [76].

Advertisement

7. Hydroxylation of aromatics with H2O2

Green hydroxylation of aromatics to phenolic derivatives, such as phenol, cresol, catechol, and hydroquinone, is still a huge challenge for producing many kinds of fine chemicals in industry, due to the strong charge delocalization effect of six‐member ring of benzene [7779]. The production of phenol is mainly on the basis of the commercial cumene process, which contains a three‐step process (that is alkylation of benzene with propylene, cumene oxidation, and acid‐catalyzed decomposition), with acetone as the major by‐product. As acetone is very cheap and superfluous than its consumption, this process is suffered by the capital cost and acetone market. Moreover, this route has a very complex and long pathway, and gives birth to a large amount of by‐products due to the severe reaction conditions. As a result, several selective catalysts were proposed, such as V, Fe, and Cu‐based catalysts, but they are far beyond to be used in a large scale due to their poor catalytic activity or stability. Direct benzene oxidation to phenol with H2O2 catalyzed by TS‐1 is widely accepted as an alternative process to the traditional cumene process for producing phenol, as shown in Scheme 5. As previously mentioned, the function of TS‐1 catalyst involves the activation of H2O2 oxidant to form a highly reactive Ti‐OOH species, which can easily nucleophilic attack the C atoms in benzene molecules. Moreover, TS‐1 zeolite is ultrastable and can be even used in the strong acid media. Some other metal‐incorporated zeolites, such as Ti‐beta, Ti‐MOR, and V‐beta, were also evaluated by direct benzene oxidation, but they were of lower catalytic performance than TS‐1.

Scheme 5.

Mechanism of benzene hydroxylation with H2O2 catalyzed by TS‐1 zeolite.

Furthermore, phenol hydroxylation to produce catechol and hydroquinone is also of great importance for fine chemical production, such as agrichemical, photography chemicals, antioxidants, polymerization inhibitors, etc [80, 81]. In conventional routes, the dihydroxybenzenes are commercially produced by several complex processes, including (i) the oxidations of aniline and diisopropylbenzene and (ii) sulfurization of benzene and hydrolysis of benzene sulfonate [82]. Table 2 shows the detailed technologies for the production of dihydroxybenzenes in industry, but some of them have been sifted out, owing to inevitable limitations. We can see that HTS zeolite demonstrates the highest catalytic performance and technical advantages among these catalysts, especially when taking the environmental and cost benefits into consideration. It was developed and commercially scaled up by EniChem in 1986 for the first time to substitute the Brichima process, avoiding the separation problem of homogeneous catalyst.

Approach Catalyst Xphenol (%) Sp‐ and o‐product (%)
Rhone‐Poulenc H3PO4/HClO4 5 90
Ube Ketone peroxide/H2SO4 <5 90
Brichima Fe(II)/Co(II) 10 80
EniChem* (H2O2/phenol = 1/3) HTS 26 >95

Table 2.

Different phenol hydroxylation approaches to produce catechol and hydroquinone.

After 2000, phenol hydroxylation catalyzed by HTS zeolite was established, with significant reduction of financial investment and operation cost. It is worthy to note that the catalytic activity of TS‐1 zeolite in this reaction is closely associated with the chemical status of Ti species; only tetrahedral framework Ti species are able to promote the activation of H2O2 oxidant, rather than the substrate. Therefore, phenol hydroxylation is usually chosen as a probe reaction to reveal the intrinsic activity of framework Ti sites incorporated.

Additionally, to illustrate the preponderance of hierarchical structure, both TS‐1 and HTS zeolites were evaluated by several aromatic (including benzene, toluene and phenol) oxidation reactions, as shown in Figure 11. It is observed that HTS zeolite exhibits much higher catalytic performance than TS‐1 zeolite in all aromatic oxidation we tested, with ultralong lifetime in these reactions as well. This ascribes to the higher content of framework Ti species of HTS than TS‐1, and the improvement of mass diffusion via the formation of hollow cavities, which is in great agreement with the dissolution‐recrystallization process through post‐synthesis treatment. Accordingly, these results suggest that phenol molecules are more active than toluene, and far more reactive than benzene, in TS‐1 catalyzed aromatic oxidations, due to the change of charge property by introducing electronic groups. It is interesting to find that the solvent effect takes great impact on the catalytic performance of TS‐1 zeolite, while acetic acid can promisingly enhance toluene conversion in the presence of H+ ions.

Figure 11.

Catalytic performance of TS‐1 and HTS in various aromatic oxidation with H2O2, reaction conditions: (A) benzene/H2O2 = 1:1, solvent free, 70°C, 400 rmp; (B) toluene/H2O2 = 1:1, acetic acid as solvent, 70°C, 400 rmp; (C) phenol/H2O2 = 3:1, acetone as solvent, 80°C, 400 rmp.

Generally, the ratio of catechol to hydroquinone is close to 1:1 in the TS‐1 catalyzed phenol hydroxylation route, and catechol (para‐dihydroxybenzene) is more expansive and of even larger requirement than the other one. Recently, we have proposed an efficient route to modify the product distribution, with obviously increasing of para‐product selectivity, by introducing acidic‐basic sites via adding MgO‐Al2O3 binary oxides. For MgO‐Al2O3 modified HTS zeolite [83], the optimized ratio of para to ortho product is larger than 2, while the phenol conversion still remains high (>16%). The high p‐dihydroxybenzene selectivity is ascribed to the steric hindrance and the synergistic effect between Lewis basic‐acidic sites of mixed oxides and tetrahedral framework Ti species in HTS zeolite. In other words, HTS zeolite provides an ideal platform to transform aromatics into phenolic derivatives under mild conditions, to make these processes gradually be much cleaner and greener.

Advertisement

8. Baeyer‐Villiger oxidation of cyclohexanone with H2O2

Baeyer‐Villiger (BV) oxidation is of great importance to convert ketones to esters or lactones, which are important organic intermediates in the fine chemical industry [8490]. ε‐carprolactone is traditionally produced from the BV oxidation reaction of cyclohexanone by using peracids as a catalyst and oxidant, e.g., perbenzoic acid, m‐chloroperbenzoic acid, and trifluoroperacetic acid. Although peracids route favors high ketone conversion and fast product transformation, it suffers serious pollution and operation drawbacks at commercial scale [9197]. Aqueous hydrogen peroxide solutions (usually with 30–50 wt% concentration), environmental friendly, cheap, and easy to handle, are considered as a desirable oxidant to replace hazardous peracids. However, hydrogen peroxide is usually of less attacking nucleophilic capability than peracids for activating the carbonyl group of ketones [98102].

In previous works, much effort has been devoted to develop both homogeneous and heterogeneous catalysts in order to improve the performance of BV oxidation of cyclohexanone [104111]. One famous catalyst is Sn‐beta zeolite, reported by A. Corma, with almost 100% ε‐carprolactone selectivity in direct BV oxidation of cyclohexanone in organic solvent at low temperature (<90°C) [112]. It is confirmed that this route bases on the activation of carbonyl group by tetrahedral framework Sn species via the acceptor‐donator of electron pairs between framework Sn sites and O atoms in cyclohexanone. In contract, we have found that HTS zeolite is also an effective catalyst for the BV oxidation of cyclohexanone with H2O2, but further hydrolysis and oxidation of lactone have been occurred with the formation of 6‐hydroxyhexanoic acid and adipic acid, respectively. As for the mechanism of this reaction, the density function theory (DFT) method was employed, which suggests that the activation of H2O2 to form highly reactive Ti‐OOH species is the driving force of BV oxidation, as shown in Figure 12. However, due to its seven‐member‐ring structure, ε‐carprolactone easily reacts with H2O to make ring‐opening reaction occurred with the formation of 6‐hydroxyhexanoic acid. Then, the hydroxyl‐substituted acid will further be oxidized by Ti‐OOH species, to give birth to adipic acid in the end, as shown in Figure 13. To support our proposed mechanism, the major products of cyclohexanone BV oxidation were analyzed by the gas chromatography-mass spectrometry method, including 6‐hydroxyhexanoic acid, adipic acid, cyclohexane‐diol, and cyclohexane‐dione. As shown in Table 3, HTS zeolite has a much higher conversion (with longer reaction time) than Sn‐beta zeolite, without adding any organic solvent. And we also observe that increasing reaction temperature and H2O content are in favor of high cyclohexanone conversion and acid yield. It is in good agreement with the DFT result which indicated that the framework Ti species display a totally novel catalytic behavior from that Sn species incorporated, attributing to the basic difference of charge property between them.

Figure 12.

One possible reaction pathway of cyclohexanone BV oxidation catalyzed by TS‐1 zeolite involving the activation of H2O2 molecules [103].

Figure 13.

Reaction network of ring opening of ε‐carprolactone and deep oxidation of 6‐hydroxyhexanoic acid to adipic acid.

Table 3.

Reaction parameters and catalytic performance of Sn‐beta and HTS in the cyclohexanone BV oxidation with H2O2.

From the viewpoint of commercial application, although ε‐carprolactone selectivity in HTS catalyzed route is far lower, it provides an ideal and green approach to produce adipic acid, overcoming the significant drawbacks of oxidation of KA oil with HNO3, and there is no organic solvent added, reducing the energy consumption for solvent recycling and purification. Most importantly, the product distribution can be altered, according to market requirement, via changing reaction parameters and processing technologies.

Advertisement

9. Catalytic oxidation of cyclohexane with H2O2 to KA oil

Cyclohexane oxidation to prepare cyclohexanone and cyclohexanol (referred as to KA oil) is always a hot spot in the organic chemical industry, contributing to the wide application area of KA oil [113115]. Currently, KA oil is mainly manufactured by the air oxidation of cyclohexane via the formation and decomposition of cyclohexyl peroxide at over 100°C in industry, using soluble cobalt carboxylate and metal‐boric acid as a catalyst, or without a catalyst, with about 80% selectivity of KA oil, as shown in Scheme 6. However, the cyclohexane conversion should be kept at ultralow level, about 4–10% for single transformation, to reducing the selectivity of undesired products, such as adipic acid and deep hydroxylated products. Apart from the harsh reaction parameters, homogeneous catalysts are usually difficult to be separated from mixture, with high power consumption. During the past centuries, huge efforts on developing benign oxidation of cyclohexane with various oxidants (including O2, H2O2, and tert-butyl hydroperoxide (TBHP) and catalysts have been devoted. Among them, HTS zeolite exhibits delightful catalytic performance, with H2O2 as an oxidant under mild conditions, as shown in Figure 14. It is apparent to illustrate that the cyclohexane conversion for HTS zeolite is as twice the time as that of TS‐1 in the same reaction conditions, owing to the intensification of the accessibility of framework Ti species through hollow cavities. Except for the excellent conversion rate, HTS zeolite is also helpful for the high selectivity of cyclohexanone, by promoting the further oxidation of cyclohexanol with H2O2.

Scheme 6.

Comparison of cyclohexane oxidation reaction network between commercial route and TS‐1 catalyzed route with H2O2.

Figure 14.

Catalytic performance of TS‐1 and HTS in cyclohexane oxidation with H2O2 (reaction conditions: H2O2/cyclohexane = 1:1, 5 wt% catalyst content, 80°C, 500 rmp stirring).

Particularly, it should be mentioned that the selectivity of KA oil is much higher than that of air‐oxidation route, over 90% under optimized conditions. Although H2O2 solution is much more expansive than air, there is no need to carry out the complex separation and only few by‐products formed, with higher atom utilization efficiency and thus meeting the principles of green chemistry. Consequently, this route provides an alternative method for KA oil production to replace the conventional route, and the scale up study is still under investigating at SINOPEC. We confirm that it is of great industrial potential, when taking the successful commercialization cases into consideration.

Advertisement

10. Epoxidation of fatty acid methyl ester with TBHP

As the fast developing of transformation of biomass to bulk and fine chemicals, green conversion processes for biomass utilization become particularly important [116118]. In this context, epoxy fatty acid ester, one of the largest‐scale chemicals derived from biodiesel, has been widely used as environment‐friendly polyvinyl chloride (PVC) plasticizer and intermediates in the production of fine chemicals, such as nonanal, nonene, and other bulky molecules used as a biolubricant. In industry, epoxy fatty acid ester is adopted by the peracid process (Prileshajev reaction, Schemes 7 and 8). However, the Prileshajev method is of low efficiency, high risk, and pollution, due to the use of inorganic acid as catalyst and hazardous peracids as an oxidant.

Scheme 7.

FAME epoxidation process and side reactions catalyzed by TS‐SP zeolite.

Scheme 8.

Reaction pathways of the Prilezhaev reaction with H2O2 in industry.

Recently, we have developed a heterogeneous FAME epoxidation process with TBHP catalyzed by titanosilicate zeolite to replace the Prilezhaev method. While due to mass transfer limitation of FAME, TS‐1 presents low activity with FAME conversion of 16.8% and a fairly low epoxy fatty acid ester at 0.5 h. In order to overcome mass transfer limitation in TS‐1, several methods have been proposed. Recently, we have exploited a hierarchical titanosilicate zeolite with spongy structure (TS‐SP) by combining pore‐embedding and post‐treatment methods. The specific surface area of TS‐SP calculated by Brunauer‐Emmett‐Teller (BET) method (Table 4) is 518 m2·g−1, comparing with TS‐1 of 435 m2·g−1, corresponding external surface area calculated by the t‐plot method is 342 m2·g−1, which is also far more than traditional TS‐1, indicating the existing of mesopores or macropores. Pore‐size distribution and cumulative pore volume calculated by a nonlocal density functional theory (NL‐DFT) model confirmed that obvious and abundant mesopores distributed in the range of 3–5 nm were formed in the zeolite. Thus, in epoxidation of FAME, the uniform mesopores promote the diffusion of FAME and TBHP molecules inside the zeolites, leading to the better accessibility of active sites and facilitating product transport outside the pores, as shown in Figure 15. As a result, the spongy zeolite shows even more remarkable catalytic activity with 91.5% conversion at 0.5 h, which is 5.4 times as that of TS‐1. What is more, FAME could be completely consumed within only 3 h, with the epoxy product selectivity more than 90%, which is far higher than that of TS‐1. To the best of our best knowledge, this result is close to the highest performance published in the literature so far.

Figure 15.

Catalytic performance of various TS‐1 zeolites in epoxidation of methyl oleate with TBHP.

Zeolites Ti/Si ratio (%) SBET (m2g−1) Sext (m2g−1) Vmicro (cm3g−1) Dmeso (nm)
TS‐1 3.58 435 8 0.181 2.5, 4.3
TS‐SP 3.51 518 342 0.077 3–5.5

Table 4.

Textural parameters and elementary composition of titanosilicate zeolites.

In conclusion, heterogeneous catalytic epoxidation of FAME is of significant importance for the environmental protection, saving energy, and improving selectivity of target product, as it avoids the use of hazardous peracids and inorganic acids. Those achievements confirm that mass transformation limitation of FAME could be solved by introducing hierarchical pores with zeolite particles, bringing significant enhancement of catalytic activity in bulk molecule oxidation reactions.

11. Conclusion and perspective

In summary, we have systematically reviewed the synthesis of hierarchical TS‐1 zeolites and their catalytic application in several oxidation processes at SINOPEC, to replace the hazard and polluted traditional routes. In the early years, HTS zeolite was synthesized by using a “dissolution‐recrystallization” treatment under hydrothermal conditions, which demonstrates much higher catalytic performance than conventional TS‐1 zeolite, due to the diffusion improvement by intracrystalline confined mesopores. In addition, post‐synthesis helps the reincorporation of Ti species into the framework matrix, thus HTS zeolite exhibits excellent activity in many commercial processes, such as propylene epoxidation, cyclohexanone ammoximation, and phenol hydroxylation. Moreover, we have already developed several interesting oxidations, such as chlorohydrination of AC with H2O2 and HCl and cyclohexanone BV oxidation with H2O2, for meeting the requirements of sustainable and green chemistry. On the other side, we also paid more attention on the exploitation of hierarchical TS‐1 zeolites, for enhancing the entrance and mass diffusion with open mesopores, by using combined silanization and post‐treatment technologies. Therefore, the application of TS‐1 zeolites has been extended to the catalytic oxidation of bulk subtracts and oxidants (TBHP and CHHP), showing far better advantages in the epoxidation of FAME with TBHP over the Prileshajev reaction with H2O2. Most importantly, after long‐term experiment and thinking, we can rationally design and synthesize hierarchical TS‐1 zeolites with different kinds of mesopores at molecular scale, according to the demands for specific reactions. Furthermore, these catalytic materials provide bright opportunities for replacing traditionally polluted oxidation processes, and many successful cases have been achieved. For further development, many novel hierarchical zeolites with different framework composition and new topological structures will be prepared, and then these materials will also find many cleaner catalytic reactions, with an evident of reducing the generation of waste and unwanted products.

Acknowledgments

This work was financially supported by the National Basic Research Program of China (973 Program, 2006CB202508) and the China Petrochemical Corporation (SINOPEC Group 20673054).

References

  1. 1. Nemeth L, Bare SR. Science and technology of framework metal‐containing zeotype catalysts. Advances in Catalysis. 2014;57:1–97
  2. 2. Reddy JS, Kumar R, Ratnasamy P. Titanium silicalite‐2: Synthesis, characterization and catalytic properties. Applied Catalysis. 1990;58(1):1–4
  3. 3. Fejes P, Nagy JB, Kovács K, et al. Synthesis of tin (IV) silicalites (MFI) and their characterization: A Mössbauer and MAS NMR spectroscopy study. Applied Catalysis A: General. 1996;145(1–2):155–84
  4. 4. Perez‐Pariente J, Martens JA, Jacobs PA. Crystallization mechanism of zeolite beta from (TEA)2O, Na2O and K2O containing aluminosilicate gels. Applied Catalysis. 1987;31(1):35–64
  5. 5. Taramasso M, Perego G, Notari B. Preparation of porous crystalline synthetic material comprised of silicon and titanium oxides: U.S. Patent 4,410,501. 1983‐10–18
  6. 6. Zhu Y, Chuah G K, Jaenicke S. Selective Meerwein‐Ponndorf‐Verley reduction of α,β‐unsaturated aldehydes over Zr‐zeolite beta. Journal of Catalysis. 2006;241(1):25–33
  7. 7. Boronat M, Corma A, Renz M. Mechanism of the Meerwein‐Ponndorf‐Verley‐Oppenauer (MPVO) redox equilibrium on Sn− and Zr−beta zeolite catalysts. Journal of Physical Chemistry B. 2006;110(42):21168–21174
  8. 8. Pérez‐Ramírez J, Groen J C, Brückner A, et al. Evolution of isomorphously substituted iron zeolites during activation: Comparison of Fe‐beta and Fe‐ZSM‐5. Journal of Catalysis. 2005;232(2):318–334
  9. 9. Patra AK, Dutta A, Pramanik M, et al. Synthesis of hierarchical mesoporous Mn–MFI zeolite nanoparticles: A unique architecture of heterogeneous catalyst for the aerobic oxidation of thiols to disulfides. ChemCatChem. 2014;6(1):220–229
  10. 10. Kim JH, Namba S, Yashima T. Para‐selectivity of metallosilicates with MFI zeolite structure. Zeolites. 1991;11(1):59–63
  11. 11. Boronat M, Concepción P, Corma A, et al. Determination of the catalytically active oxidation Lewis acid sites in Sn‐beta zeolites, and their optimisation by the combination of theoretical and experimental studies. Journal of Catalysis. 2005;234(1):111–118
  12. 12. Van der Waal JC, Lin P, Rigutto MS, et al. Synthesis of aluminium free titanium silicate with the BEA structure using a new and selective template and its use as a catalyst in epoxidations. Studies in Surface Science and Catalysis. 1997;105:1093–1100
  13. 13. Wolf P, Hammond C, Conrad S, et al. Post‐synthetic preparation of Sn‐, Ti‐ and Zr‐beta: A facile route to water tolerant, highly active Lewis acidic zeolites. Dalton Transactions. 2014;43(11):4514–4519
  14. 14. Zhong W, Mao L, Yi W, et al. Highly efficient light‐driven HNO3 nitration‐oxidation of cyclohexane to co‐product nitrocyclohexane and adipic acid in a biphasic system. Research on Chemical Intermediates. 2016;42(2):461–470
  15. 15. Hassall CH. The Baeyer‐Villiger oxidation of aldehydes and ketones. Organic Reactions. 2011, 9(3):73–106
  16. 16. Van de Vyver S, Román‐Leshkov Y. Emerging catalytic processes for the production of adipic acid. Catalysis Science & Technology. 2013;3(6):1465–1479
  17. 17. Román‐Leshkov Y, Moliner M, Labinger JA, et al. Mechanism of glucose isomerization using a solid Lewis acid catalyst in water. Angewandte Chemie International Edition. 2010;49(47):8954–8957
  18. 18. Wu P, Tatsumi T, Komatsu T, et al. Postsynthesis, characterization, and catalytic properties in alkene epoxidation of hydrothermally stable mesoporous Ti‐SBA‐15. Chemistry of Materials. 2002;14(4):1657–1664
  19. 19. Luo HY, Lewis JD, Román‐Leshkov Y. Lewis acid zeolites for biomass conversion: Perspectives and challenges on reactivity, synthesis, and stability. Annual Review of Chemical and Biomolecular Engineering. 2016;7:663–692
  20. 20. Cesana A, Mantegazza MA, Pastori M. A study of the organic by‐products in the cyclohexanone ammoximation. Journal of Molecular Catalysis A: Chemical. 1997;117(1):367–373
  21. 21. Fornasari G, Trifiro F. Oxidation with no‐redox oxides: Ammoximation of cyclohexanone on amorphous silicas. Catalysis Today. 1998;41(4):443–455
  22. 22. Reddy JS, Sivasanker S, Ratnasamy P. Ammoximation of cyclohexanone over a titanium silicate molecular sieve, TS‐2. Journal of Molecular Catalysis. 1991;69(3):383–392
  23. 23. Bellussi G, Millini R, Pollesel P, et al. Zeolite science and technology at Eni. New Journal of Chemistry. 2016;40(5):4061–4077
  24. 24. Perego C, Carati A, Ingallina P, et al. Production of titanium containing molecular sieves and their application in catalysis. Applied Catalysis A: General. 2001;221(1): 63–72
  25. 25. Zhou J, Hua Z, Cui X, et al. Hierarchical mesoporous TS‐1 zeolite: A highly active and extraordinarily stable catalyst for the selective oxidation of 2,3,6‐trimethylphenol. Chemical Communications. 2010;46(27):4994–4996
  26. 26. Sanz R, Serrano DP, Pizarro P, et al. Hierarchical TS‐1 zeolite synthesized from SiO2‐TiO2 xerogels imprinted with silanized protozeolitic units. Chemical Engineering Journal. 2011;171(3):1428–1438
  27. 27. Wang W, Li G, Li W, et al. Synthesis of hierarchical TS‐1 by caramel templating. Chemical Communications. 2011;47(12):3529–3531
  28. 28. Wang Y, Lin M, Tuel A. Hollow TS‐1 crystals formed via a dissolution–recrystallization process. Microporous and Mesoporous Materials. 2007;102(1):80–85
  29. 29. Lin M, Shu X, Wang X. Study on the synthesis and application of environmentally friendly catalyst TS‐1 zeolite. Petroleum Processing and Petrochemicals. 1999;30:1–4
  30. 30. Xie Z, Liu Z, Wang Y, et al. Applied catalysis for sustainable development of chemical industry in China. National Science Review. 2015;2(2):167–182
  31. 31. Peng X, Xia C, Lin M, et al. Chlorohydrination of allyl chloride with HCl and H2O2 catalyzed by hollow titanium silicate zeolite to produce dichloropropanol. Green Chemistry. 2017;19:1221–1225
  32. 32. Changjiu X, Bin Z, Min L, et al. A “Green” cyclohexanone oxidation route catalyzed by hollow titanium silicate zeolite for preparing ε‐caprolactone, 6‐hydroxyhexanoic acid and adipic acid. China Petroleum Processing & Petrochemical Technology. 2012;14(4):33–41
  33. 33. Shi C, Zhu B, Lin M, et al. Propylene direct epoxidation performance of bi‐functional titanosilicate with hollow structure. Catalysis Letters. 2009;133(1–2):70
  34. 34. Lin M, Shu X, Wang X, et al. Titanium‐silicalite molecular sieve and the method for its preparation: U.S. Patent 6,475,465. 2002‐11–5
  35. 35. Wang B, Lin M, Zhu B, et al. The synthesis, characterization and catalytic activity of the hierarchical TS‐1 with the intracrystalline voids and grooves. Catalysis Communications. 2016;75:69–73
  36. 36. Wang B, Lin M, Peng X, et al. Hierarchical TS‐1 synthesized effectively by post‐modification with TPAOH and ammonium hydroxide. RSC Advances. 2016;6(51):44963–44971
  37. 37. Xia C, M Lin, B Zhu, Y Luo, X Shu. Hollow titanium silicalite zeolite: From fundamental research to commercial application in environmental‐friendly catalytic oxidation processes. In: Zeolites-Useful Minerals. InTech, Croatia; 2016
  38. 38. Serrano DP, Pizarro P. Synthesis strategies in the search for hierarchical zeolites. Chemical Society Reviews. 2013;42(9):4004–4035
  39. 39. Serrano DP, Sanz R, Pizarro P, et al. Tailoring the properties of hierarchical TS‐1 zeolite synthesized from silanized protozeolitic units. Applied Catalysis A: General. 2012;435:32–42
  40. 40. Bordiga S, Bonino F, Damin A, et al. Reactivity of Ti (iv) species hosted in TS‐1 towards H2O2–H2O solutions investigated by ab initio cluster and periodic approaches combined with experimental XANES and EXAFS data: A review and new highlights. Physical Chemistry Chemical Physics. 2007;9(35):4854–4878
  41. 41. Guo Q, Sun K, Feng Z, et al. A thorough investigation of the active titanium species in TS‐1 zeolite by in situ UV resonance Raman spectroscopy. Chemistry: A European Journal. 2012;18(43):13854–13860
  42. 42. Zhang J, Tan J, Wang K, et al. Chlorohydrination of allyl chloride to dichloropropanol in a microchemical system. Industrial & Engineering Chemistry Research. 2012;51(45):14685–14691
  43. 43. Machado RM, Chapman TW. Kinetics and selectivity of aqueous propylene halohydrination. Industrial & Engineering Chemistry Research. 1989;28(12):1789–1794
  44. 44. Luo ZH, You XZ, Li HR. Direct preparation kinetics of 1,3‐dichloro‐2‐propanol from glycerol using acetic acid catalyst. Industrial & Engineering Chemistry Research. 2008;48(1):446–452
  45. 45. Gao H, Lu G, Suo J, et al. Epoxidation of allyl chloride with hydrogen peroxide catalyzed by titanium silicalite 1. Applied Catalysis A: General. 1996;138(1):27–38
  46. 46. Wang Q, Mi Z, Wang Y, et al. Epoxidation of allyl choride with molecular oxygen and 2‐ethyl‐anthrahydroquinone catalyzed by TS‐1. Journal of Molecular Catalysis A: Chemical. 2005;229(1):71–75
  47. 47. Russo V, Tesser R, Santacesaria E, et al. Chemical and technical aspects of propene oxide production via hydrogen peroxide (HPPO process). Industrial & Engineering Chemistry Research. 2013;52(3):1168–1178
  48. 48. Xiong G, Cao Y, Guo Z, et al. The roles of different titanium species in TS‐1 zeolite in propylene epoxidation studied by in situ UV Raman spectroscopy. Physical Chemistry Chemical Physics. 2016;18(1):190–196
  49. 49. Lu Y, Wang R, Zhang J, et al. Evaluation of an improved epichlorohydrin synthesis from dichloropropanol using a microchemical system. Chinese Journal of Chemical Engineering. 2015;23(7):1123–1130
  50. 50. Nakagawara S, Ozawa Y, Hotta K, et al. Spectroscopic characterization and the pH dependence of bactericidal activity of the aqueous chlorine solution. Analytical Sciences. 1998;14(4):691–698
  51. 51. Yoon BH, Wang LJ, Yoon SL, et al. Mechanism of chlorate formation in chlorine dioxide delignification. Appita Journal: Journal of the Technical Association of the Australian and New Zealand Pulp and Paper Industry. 2004;57(6):472
  52. 52. Geiger F M, Hicks JM, de Dios AC. Ab initio study of HOCl, HCl, H2O, and Cl2 interacting with four water molecules. The Journal of Physical Chemistry A. 1998;102(9):1514–1522
  53. 53. Wu P, Komatsu T, Yashima T. Ammoximation of ketones over titanium mordenite. Journal of Catalysis. 1997;168(2):400–411
  54. 54. Thangaraj A, Sivasanker S, Ratnasamy P. Catalytic properties of crystalline titanium silicalites III. Ammoximation of cyclohexanone. Journal of Catalysis. 1991;131(2):394–400
  55. 55. Song F, Liu Y, Wang L, et al. Highly selective synthesis of methyl ethyl ketone oxime through ammoximation over Ti‐MWW. Applied Catalysis A: General. 2007;327(1):22–31
  56. 56. Chu CQ, Zhao HT, Qi YY, et al. Density functional theory studies on hydroxylamine mechanism of cyclohexanone ammoximation on titanium silicalite‐1 catalyst. Journal of Molecular Modeling. 2013;19(6):2217–2224
  57. 57. Zecchina A, Spoto G, Bordiga S, et al. Ammoximation of cyclohexanone on titanium silicalite: Investigation of the reaction mechanism. Studies in Surface Science and Catalysis. 1993;75:719–729
  58. 58. Petrini G, Leofanti G, Mantegazza MA, et al. Caprolactam via ammoximation. 1996.
  59. 59. Mantegazza MA, Petrini G, Spano G, et al. Selective oxidations with hydrogen peroxide and titanium silicalite catalyst. Journal of Molecular Catalysis A: Chemical. 1999;146(1):223–228
  60. 60. Xuhong M, Dianzhong W, Yongrui W, et al. Nanosized molecular sieves as petroleum refining and petrochemical catalysts. Chinese Journal of Catalysis. 2013;34(1):69–79
  61. 61. Wang Y, Zhang S, Zhao Y, et al. The mechanism of catalyst deactivation and by‐product formation in acetone ammoximation catalyzed by hollow titanium silicalite. Journal of Molecular Catalysis A: Chemical. 2014;385:1–6
  62. 62. Munakata H, Oumi Y, Miyamoto A. A DFT study on peroxo‐complex in titanosilicate catalyst: Hydrogen peroxide activation on titanosilicalite‐1 catalyst and reaction mechanisms for catalytic olefin epoxidation and for hydroxylamine formation from ammonia. The Journal of Physical Chemistry B. 2001;105(17):3493–3501
  63. 63. Zecchina A, Bordiga S, Lamberti C, et al. Structural characterization of Ti centres in Ti‐silicalite and reaction mechanisms in cyclohexanone ammoximation. Catalysis Today. 1996;32(1–4):97–106
  64. 64. Xia C, Lin M, Zheng A, et al. Irreversible deactivation of hollow TS‐1 zeolite caused by the formation of acidic amorphous TiO2–SiO2 nanoparticles in a commercial cyclohexanone ammoximation process. Journal of Catalysis. 2016;338:340–348
  65. 65. Xia C, Lin M, Peng X, et al. Regeneration of deactivated hollow titanium silicalite zeolite from commercial ammoximation process by encapsulating amorphous TiO2‐SiO2 nanoparticles inside zeolite crystal. Chemistry Select. 2016;1(14):4187–4192
  66. 66. Zheng A, Xia C, Xiang Y, et al. Titanium species in deactivated HTS‐1 zeolite from industrial cyclohexanone ammoxidation process. Catalysis Communications. 2014;45:34–38
  67. 67. Meiers R, Hölderich WF. Epoxidation of propylene and direct synthesis of hydrogen peroxide by hydrogen and oxygen. Catalysis Letters. 1999;59(2):161–163
  68. 68. Cumaranatunge L, Delgass WN. Enhancement of Au capture efficiency and activity of Au/TS‐1 catalysts for propylene epoxidation. Journal of Catalysis. 2005;232(1):38–42
  69. 69. Serrano DP, Sanz R, Pizarro P, et al. Preparation of extruded catalysts based on TS‐1 zeolite for their application in propylene epoxidation. Catalysis Today. 2009;143(1):151–157
  70. 70. Sharifov GS. Kinetics of chlorohydrination of allyl chloride. Kinetics and Catalysis. 2001;42(5):615–619
  71. 71. Burgin J, Hearne G, Rust F. Derivatives of allylic chlorides. Industrial & Engineering Chemistry. 1941;33(3):385–388
  72. 72. Kiryu T, Abe K, Honma A, et al. Chlorohydrination of allylchloride and methallylchloride. Journal of the Japan Petroleum Institute. 1969;12(5):350–355
  73. 73. http://corporate.evonik.com/en/company/locations/location‐asia/pages/skc‐evonik‐peroxide‐korea‐co‐ltd.aspx
  74. 74. http://www.chemicals‐technology.com/projects/basf‐hppo/
  75. 75. Bassler P, Weidenbach M, Goebbel H. Chemical Engineering Transactions. 2010;21:571–576
  76. 76. Lin M, Xia C, Zhu B, et al. Green and efficient epoxidation of propylene with hydrogen peroxide (HPPO process) catalyzed by hollow TS‐1 zeolite: A 1.0 kt/a pilot‐scale study. Chemical Engineering Journal. 2016;295:370–375
  77. 77. Tanev PT, Chibwe M, Pinnavaia TJ. Titanium‐containing mesoporous molecular sieves for catalytic oxidation of aromatic compounds. Nature. 1994;368(6469):321
  78. 78. Mimoun H, Saussine L, Daire E, et al. Vanadium (V) peroxy complexes. New versatile biomimetic reagents for epoxidation of olefins and hydroxylation of alkanes and aromatic hydrocarbons. Journal of the American Chemical Society. 1983;105(10):3101–3110
  79. 79. Thangaraj A, Kumar R, Ratnasamy P. Catalytic properties of crystalline titanium silicalites II. Hydroxylation of phenol with hydrogen peroxide over TS‐1 zeolites. Journal of Catalysis. 1991;131(1):294–297
  80. 80. Varagnat J. Hydroquinone and pyrocatechol production by direct oxidation of phenol. Industrial & Engineering Chemistry Product Research and Development. 1976;15(3):212–215
  81. 81. Neidig ML, Hirsekorn KF. Insight into contributions to phenol selectivity in the solution oxidation of benzene to phenol with H2O2. Catalysis Communications. 2011;12(6):480–484
  82. 82. Suresh S, Srivastava VC, Mishra IM. Adsorption of catechol, resorcinol, hydroquinone, and their derivatives: A review. International Journal of Energy and Environmental Engineering. 2012;3(1):32
  83. 83. Xia C, Long L, Zhu B, et al. Enhancing the selectivity of para‐dihydroxybenzene in hollow titanium silicalite zeolite catalyzed phenol hydroxylation by introducing acid–base sites. Catalysis Communications. 2016;80:49–52
  84. 84. Jiménez‐Sanchidrián C, Ruiz JR. The Baeyer–Villiger reaction on heterogeneous catalysts. Tetrahedron. 2008;64(9):2011–2026
  85. 85. Ten Brink GJ, Arends I, Sheldon RA. The Baeyer‐Villiger reaction: New developments toward greener procedures. Chemical Reviews. 2004;104(9):4105–4124
  86. 86. Krow GR. The Baeyer‐Villiger oxidation of ketones and aldehydes. Organic Reactions. 2004;43(3):251–798
  87. 87. Strukul G. Transition metal catalysis in the Baeyer‐Villiger oxidation of ketones. Angewandte Chemie International Edition. 1998;37(9):1198–1209
  88. 88. Kotlewska AJ, van Rantwijk F, Sheldon RA, et al. Epoxidation and Baeyer‐Villiger oxidation using hydrogen peroxide and a lipase dissolved in ionic liquids. Green Chemistry. 2011;13(8):2154–2160
  89. 89. Figg TM, Webb JR, Cundari TR, et al. Carbon‐oxygen bond formation via organometallic Baeyer‐Villiger transformations: A computational study on the impact of metal identity. Journal of the American Chemical Society. 2012;134(4):2332–2339
  90. 90. Renz M, Meunier B. 100 years of Baeyer‐Villiger oxidations. European Journal of Organic Chemistry. 1999;1999(4):737–750
  91. 91. Pande CS, Jain N. Polymer‐supported persulfonic acid as oxidising agent. Synthetic Communications. 1989;19(7–8):1271–1279
  92. 92. Renz M, Blasco T, Corma A, et al. Selective and shape‐selective Baeyer‐Villiger oxidations of aromatic aldehydes and cyclic ketones with Sn‐beta zeolites and H2O2. Chemistry: A European Journal. 2002;8(20):4708–4717
  93. 93. Strukul G, Varagnolo A, Pinna F. New (old) hydroxo complexes of platinum (II) as catalysts for the Baeyer‐Villiger oxidation of ketones with hydrogen peroxide. Journal of Molecular Catalysis A: Chemical. 1997;117(1–3):413–423
  94. 94. Corma A, Llabrés i Xamena FX, Prestipino C, et al. Water resistant, catalytically active Nb and Ta isolated Lewis acid sites, homogeneously distributed by direct synthesis in a beta zeolite. The Journal of Physical Chemistry C. 2009;113(26):11306–11315
  95. 95. Canan Koch SS, Chamberlin AR. Modified conditions for efficient Baeyer‐Villiger oxidation with m‐CPBA. Synthetic Communications. 1989;19(5–6):829–833
  96. 96. Bernini R, Coratti A, Fabrizi G, et al. CH3ReO3/H2O2 in room temperature ionic liquids: An homogeneous recyclable catalytic system for the Baeyer‐Villiger reaction. Tetrahedron Letters. 2003;44(50):8991–8994
  97. 97. Kaneda K, Yamashita T. Heterogeneous Baeyer‐Villiger oxidation of ketones using m‐chloroperbenzoic acid catalyzed by hydrotalcites. Tetrahedron Letters. 1996;37(26):4555–4558
  98. 98. Bolm C, Schlingloff G, Weickhardt K. Optically active lactones from a Baeyer‐Villiger‐type metal‐catalyzed oxidation with molecular oxygen. Angewandte Chemie International Edition (in English). 1994;33(18):1848–1849
  99. 99. Watanabe A, Uchida T, Ito K, et al. Highly enantioselective Baeyer–Villiger oxidation using Zr (salen) complex as catalyst. Tetrahedron Letters. 2002;43(25):4481–4485
  100. 100. Punniyamurthy T, Velusamy S, Iqbal J. Recent advances in transition metal catalyzed oxidation of organic substrates with molecular oxygen. Chemical Reviews. 2005;105(6):2329–2364
  101. 101. Zhu Z, Xu H, Jiang J, et al. Postsynthesis of FAU‐type stannosilicate as efficient heterogeneous catalyst for Baeyer‐Villiger oxidation. Applied Catalysis A: General. 2016;519:155–164
  102. 102. Martins LM, Pombeiro AJL. C‐scorpionate rhenium complexes and their application as catalysts in Baeyer‐Villiger oxidation of ketones. Inorganica Chimica Acta. 2017;455:390–397
  103. 103. Xia C, Ju L, Zhao Y, et al. Heterogeneous oxidation of cyclohexanone catalyzed by TS‐1: Combined experimental and DFT studies. Chinese Journal of Catalysis. 2015;36(6):845–854
  104. 104. Kaneda K, Ueno S, Imanaka T. Heterogeneous Baeyer‐Villiger oxidation of ketones using an oxidant consisting of molecular oxygen and aldehydes in the presence of hydrotalcite catalysts. Journal of the Chemical Society, Chemical Communications. 1994;(7):797–798
  105. 105. Pillai U R, Sahle‐Demessie E. Sn‐exchanged hydrotalcites as catalysts for clean and selective Baeyer‐Villiger oxidation of ketones using hydrogen peroxide. Journal of Molecular Catalysis A: Chemical. 2003;191(1):93–100
  106. 106. Lei Z, Zhang Q, Luo J, et al. Baeyer‐Villiger oxidation of ketones with hydrogen peroxide catalyzed by Sn‐palygorskite. Tetrahedron Letters. 2005;46(20):3505–3508
  107. 107. Ueno S, Ebitani K, Ookubo A, et al. The active sites in the heterogeneous Baeyer‐Villiger oxidation of cyclopentanone by hydrotalcite catalysts. Applied Surface Science. 1997;121:366–371
  108. 108. Kawabata T, Ohishi Y, Itsuki S, et al. Iron‐containing MCM‐41 catalysts for Baeyer‐Villiger oxidation of ketones using molecular oxygen and benzaldehyde. Journal of Molecular Catalysis A: Chemical. 2005;236(1):99–106
  109. 109. Li J, Le Y, Dai WL, et al. Self‐assembled Mg5(CO3)4(OH)2·4H2O nanosheet as an effective catalyst in the Baeyer‐Villiger oxidation of cyclohexanone. Catalysis Communications. 2008;9(6):1334–1341
  110. 110. Raja R, Thomas JM, Sankar G. Baeyer‐Villiger oxidations with a difference: Molecular sieve redox catalysts for the low‐temperature conversion of ketones to lactones. Chemical Communications. 1999;(6):525–526
  111. 111. Rafelt JS, Clark JH. Recent advances in the partial oxidation of organic molecules using heterogeneous catalysis. Catalysis Today. 2000;57(1):33–44
  112. 112. Corma A, Nemeth L T, Renz M, et al. Sn‐zeolite beta as a heterogeneous chemoselective catalyst for Baeyer‐Villiger oxidations. Nature. 2001;412(6845):423–425
  113. 113. Schuchardt U, Cardoso D, Sercheli R, et al. Cyclohexane oxidation continues to be a challenge. Applied Catalysis A: General. 2001;211(1):1–17
  114. 114. Raja R, Sankar G, Thomas JM. Powerful redox molecular sieve catalysts for the selective oxidation of cyclohexane in air. Journal of the American Chemical Society. 1999;121(50):11926–11927
  115. 115. Perkas N, Wang Y, Koltypin Y, et al. Mesoporous iron‐titania catalyst for cyclohexane oxidation. Chemical Communications. 2001;(11):988–989
  116. 116. Guidotti M, Ravasio N, Psaro R, et al. Epoxidation of unsaturated FAMEs obtained from vegetable source over Ti (IV)‐grafted silica catalysts: A comparison between ordered and non‐ordered mesoporous materials. Journal of Molecular Catalysis A: Chemical. 2006;250(1):218–225
  117. 117. Ucciani E, Debal A, Rafaralahitsimba G. Epoxidation of fatty acid methyl esters with organic hydroperoxides and molybdenum oxide. European Journal of Lipid Science and Technology. 1993;95(6):236–239
  118. 118. Wilde N, Pelz M, Gebhardt S G, et al. Highly efficient nano‐sized TS‐1 with micro‐/mesoporosity from desilication and recrystallization for the epoxidation of biodiesel with H2O2. Green Chemistry. 2015;17(6):3378–3389

Written By

Changjiu Xia, Xinxin Peng, Yao Zhang, Baorong Wang, Min Lin, Bin Zhu, Yibin Luo and Xingtian Shu

Submitted: 08 November 2016 Reviewed: 08 March 2017 Published: 05 July 2017